{ "9804/astro-ph9804073_arXiv.txt": { "abstract": "We introduce a statistical quantity, known as the $K$ function, related to the integral of the two--point correlation function. It gives us straightforward information about the scale where clustering dominates and the scale at which homogeneity is reached. We evaluate the correlation dimension, $D_2$, as the local slope of the log--log plot of the $K$ function. We apply this statistic to several stochastic point fields, to three numerical simulations describing the distribution of clusters and finally to real galaxy redshift surveys. Four different galaxy catalogues have been analysed using this technique: the Center for Astrophysics I, the Perseus--Pisces redshift surveys (these two lying in our local neighbourhood), the Stromlo--APM and the 1.2 Jy {\\it IRAS} redshift surveys (these two encompassing a larger volume). In all cases, this cumulant quantity shows the fingerprint of the transition to homogeneity. The reliability of the estimates is clearly demonstrated by the results from controllable point sets, such as the segment Cox processes. In the cluster distribution models, as well as in the real galaxy catalogues, we never see long plateaus when plotting $D_2$ as a function of the scale, leaving no hope for unbounded fractal distributions. ", "introduction": "The standard cosmology is based on the assumption that the Universe must be homogeneous on very large scales. Several pieces of evidence support this assumption: the homogeneity and isotropy of the microwave background radiation \\cite{cobe} and some aspects of the large scale distribution of matter \\cite{peb89} seem to strongly advocate uniformity on scales bigger than about $200 \\, h^{-1}$ Mpc (where $H_0= 100 h$ km s$^{-1}$ Mpc$^{-1}$). However the presence of very large features in the galaxy distribution like the Bootes void \\cite{kir81} or the Great Wall \\cite{gel89} which span a scale length of the order of $100 \\, h^{-1}$ Mpc calls the actual scale of homogeneity into question. Moreover other authors consider the assumption of homogeneity just a theoretical prejudice not necessarily supported by the observational evidence quoted above. They defend the alternative idea of an unbounded fractal cosmology \\cite{col92}. Guzzo (1997) argues against this interpretation on the basis of a careful handling of the data. The spatial two--point correlation function is the statistical tool mainly used to describe the clustering in the Universe (Peebles 1980, 1993). However, because of the integral constraint \\cite{p80}, one cannot estimate it at very large distances from the currently available redshift surveys. In order to study clustering in the regime where it is not very strong, we have only two possibilities: either we extend the size of the redshift catalogues or we use alternative statistical descriptors. The approach described in this paper points in the latter direction. In the same line, other authors \\cite{fis93,par94,tad96} have tried to measure the power--spectrum on large scales directly from galaxy catalogues. Einasto \\& Gramman (1993) studied the transition to homogeneity by means of the power--spectrum and found a relation between the correlation transition scale and the spectral transition scale (turnover in $P(k)$). We introduce the quantity called $K(r)$, which is related to the correlation function $\\xi(r)$. The novelty of our approach lies essentially in the fact that we shall use a cumulant quantity instead of a differential quantity such as $\\xi(r)$. Although for a point process the functions $\\xi(r)$ and $K(r)$ are well defined, what we measure from the galaxy catalogues are just estimators of those functions. One of our main claims is that the estimators for $K(r)$ are more reliable than the most currently used estimators for $\\xi(r)$ and that makes its use recommendable (especially in three-dimensional processes and at large scales) despite its somewhat less informative character. ", "conclusions": "We should like also to comment briefly on the relation of $K$ with the correlation function $\\xi(r)$. Both play their role in the analysis of the point pattern and, as Stoyan \\& Stoyan (1996) say, their relation is similar to that between the distribution function and the probability density function in classical statistics. The use of a cumulative quantity such as $K$ avoids binning in distance, which is often a source of arbitrariness for $\\xi$ \\cite{rip92}. Let us explain why $\\xi$ does suffer from the hindrance of splitting the information into disjoint bins. When one estimates $\\xi(r)$ in $[r,r+dr]$, it is assumed that within that bin the correlation function is constant, and since this is obviously not true, the larger the bin the larger the error, but we cannot make arbitrarily small the size $dr$ of the bin, because in that case we would not find any pairs. In other words, $\\xi(r)$ has an additional source of bias, not present in $K$, due to the smoothing caused by averaging over pairs of points close to but not exactly $r$ units apart of each other (Stein 1996). The correlation length ($r_0| \\xi(r_0)=1$) is just the scale at which the density of galaxies is, on average, twice the mean number density. At smaller scales the pair correlations are due to non--linear perturbations, but homogeneity is not reached till $\\xi(r_{\\rm \\tiny hom}) \\sim 0$. The main interest of $K$ is that it permits us to study clustering precisely in that \\lq\\lq difficult'' range where $r_0 < r < r_{\\rm \\tiny hom}$, which cannot be reached by $\\xi$ because in this range the errors on the estimates of $\\xi$ are comparable with their values, while the difference $K-K_{\\rm \\tiny Pois}$ is still meaningful. As a concluding remark, we want to stress that an unbiased estimator of a quantity related with the correlation integral, known as the $K$ function, has been applied to cosmological simulations and galaxy samples. This function, extensively used in the field of spatial statistics, provides a nice measure of clustering. The border correction used here does not waste any data points and does not introduce spurious homogeneization, giving reliability to the evaluation of this function at large scales. Through the slope of $K$ we are able to calculate $D_2$, which is an indicator of a possible fractal behaviour of the point process at a given scale range. The clear physical meaning of $K$ and $D_2$ helps us easily interpret the clustering properties of different models of structure formation at different scales. Regarding the analysis of the galaxy redshift surveys, we have seen that the estimator of the $K$ function is robust in the sense that it does not depend on the shape of the study region and provides us with reliable information about the point patterns over a wide range of scales. The behaviour of the local dimension $D_2$ for the real galaxy samples is particularly interesting to proponents of various fractal models of large--scale structure. If a constancy of $D_2$ with the scale is a necessary condition for having a fractal point pattern (although it should not be sufficient as we have seen with the Cox process [see also Stoyan (1994) for more examples]), it is a neat conclusion of our analysis that the galaxy distribution does not even hold the necessary condition. The analysis presented here will provide a conclusive test to discover the scale at which the distribution of the matter in the Universe is really homogeneous when applied, in the near future, to the bigger and deeper galaxy catalogues which will be soon ready for common use. \\subsection*{ACKNOWLEDGEMENTS} This work has been partially supported by an EC Human Capital and Mobility network (contract ERB CHRX-CT93-0129) and by the Spanish DGES project n. PB96-0707. We thank prof. Stoyan for bringing the Cox model to our attention and for useful conversations and comments. We thank R. Croft, S. Paredes, R. Trasarti--Battistoni and R. van de Weygaert for kindly allowing us to use their samples and programs, as well as T. Buchert, J. Schmalzing, M. Stein and specially M. Kerscher for very interesting discussions and comments. The authors want to thank the anonymous referee for his/her valuable comments and suggestions." }, "9804/gr-qc9804034_arXiv.txt": { "abstract": "Primordial black holes may form in the early Universe, for example from the collapse of large amplitude density perturbations predicted in some inflationary models. Light black holes undergo Hawking evaporation, the energy injection from which is constrained both at the epoch of nucleosynthesis and at the present. The failure as yet to unambiguously detect primordial black holes places important constraints. In this article, we are particularly concerned with the dependence of these constraints on the model for the complete cosmological history, from the time of formation to the present. Black holes presently give the strongest constraint on the spectral index $n$ of density perturbations, though this constraint does require $n$ to be constant over a very wide range of scales. ", "introduction": "Black holes are tenacious objects, and any which form in the very early Universe are able to survive until the present, unless their Hawking evaporation is important. The lifetime of an evaporating black hole is given by \\begin{equation} \\frac{\\tau}{10^{17} \\, {\\rm sec}} \\simeq \\left( \\frac{M}{10^{15} \\, {\\rm grams}} \\right)^3 \\,. \\end{equation} From this we learn that a black hole of initial mass $M \\sim 10^{15}$g will evaporate at the present epoch, while for significantly heavier black holes Hawking evaporation is negligible. Another mass worthy of consideration is $M \\sim 10^{9}$g, which leads to evaporation around the time of nucleosynthesis, which is well enough understood to tolerate only modest interference from black hole evaporation by-products. Several mechanisms have been proposed which might lead to black hole formation; the simplest is collapse from large-amplitude, short-wavelength density perturbations. They will form with approximately the horizon mass, which in a radiation-dominated era is given by \\begin{equation} \\label{hormass} M_{{\\rm HOR}} \\simeq 10^{18} \\, {\\rm g} \\, \\left( \\frac{10^7 \\, {\\rm GeV}}{T} \\right)^2 \\,, \\end{equation} where $T$ is the ambient temperature. This tells us that any black holes for which evaporation is important must have formed during very early stages of the Universe's evolution. In particular, formation corresponds to times much earlier than nucleosynthesis (energy scale of abut $1\\,$MeV), which is the earliest time that we have any secure knowledge concerning the evolution of the Universe. Any modelling of the evolution of the Universe before one second is speculative, and especially above the electro-weak symmetry breaking scale (about $100 \\,$GeV) many possibilities exist. Note also that although we believe we understand the relevant physics up to the electro-weak scale, the cosmology between that scale and nucleosynthesis could still be modified, say by some massive but long-lived particle. In this article we will consider the standard cosmology and two alternatives \\cite{gl,glr}. We define the mass fraction of black holes as \\begin{equation} \\beta \\equiv \\frac{\\rho_{{\\rm pbh}}}{\\rho_{{\\rm tot}}} \\,, \\end{equation} and will use subscript `i' to denote the initial values. In fact, we will normally prefer to use \\begin{equation} \\alpha \\equiv \\frac{\\rho_{{\\rm pbh}}}{\\rho_{{\\rm tot}}-\\rho_{{\\rm pbh}}} = \\frac{\\beta}{1-\\beta} \\,, \\end{equation} which is the ratio of the black hole energy density to the energy density of everything else. Black holes typically offer very strong constraints because after formation the black hole energy density redshifts away as non-relativistic matter (apart from the extra losses through evaporation). In the standard cosmology the Universe is radiation dominated at these times, and so the energy density in black holes grows, relative to the total, proportionally to the scale factor $a$. As interesting black holes form so early, this factor can be extremely large, and so typically the initial black hole mass fraction is constrained to be very small. The constraints on evaporating black holes are well known, and we summarize them in Table~\\ref{massfrac}. This table shows the allowed mass fractions at the time of evaporation. An additional, optional, constraint can be imposed if one imagines that black hole evaporation leaves a relic particle, as these relics must then not over-dominate the mass density of the present Universe \\cite{BCL:PBH}. For black holes massive enough to have negligible evaporation, the mass density constraint is the only important one (though in certain mass ranges there are also microlensing limits which are somewhat stronger). \\begin{table}[t] \\caption[massfrac]{\\label{massfrac} Limits on the mass fraction of black holes at evaporation.} \\begin{tabular}{|c|c|c|} \\hline \\hline Constraint & Range & Reason \\\\ \\hline $\\alpha_{\\rm{evap}} < 0.04$ & $10^{9}$ g $< M < 10^{13}$ g & Entropy per baryon\\\\ & & at nucleosynthesis \\cite{var:ent} \\\\ \\hline $\\alpha_{\\rm{evap}} < 10^{-26} \\frac{M}{m_{{\\rm Pl}}}$ & $M \\simeq 5\\times10^{14}$~g & $\\gamma$ rays from current\\\\ & & explosions \\cite{var:gam} \\\\ \\hline $\\alpha_{\\rm{evap}} < 6\\times10^{-10} \\left( \\frac{M}{m_{{\\rm Pl}}}\\right)^{1/2}$ & $10^{9}$~g $ < M <10^{11}$~g & n$\\bar{\\rm{n}}$ production \\\\ & & at nucleosynthesis \\cite{var:neu} \\\\ \\hline $\\alpha_{\\rm{evap}} < 5\\times10^{-29} \\left( \\frac{M}{m_{{\\rm Pl}}}\\right)^{3/2}$ & $10^{10}$~g $< M < 10^{11}$~g & Deuterium destruction \\cite{lin:deu} \\\\ \\hline $\\alpha_{\\rm{evap}} < 1\\times10^{-59}\\left( \\frac{M}{m_{{\\rm Pl}}}\\right)^{7/2}$ & $10^{11}$~g $< M < 10^{13}$~g & Helium-4 spallation \\cite{var:he4}\\\\ \\hline \\end{tabular} \\vspace*{2pc} \\end{table} We will study three different cosmological histories in this paper, all of which are currently observationally viable. The first, which we call the standard cosmology, is the minimal scenario. It begins at some early time with cosmological inflation, which is necessary in order to produce the density perturbations which will later collapse to form black holes. Inflation ends, through the preheating/reheating transition (which we will take to be short), giving way to a period of radiation domination. Radiation domination is essential when the Universe is one second old, in order for successful nucleosynthesis to proceed. Finally, radiation domination gives way to matter domination, at a redshift $z_{{\\rm eq}} = 24\\,000\\,\\Omega_0 h^2$ where $\\Omega_0$ and $h$ have their usual meanings, to give our present Universe. The two modified scenarios replace part of the long radiation-dominated era between the end of inflation and nucleosynthesis. The first possibility is that there is a brief second period of inflation, known as thermal inflation \\cite{ls}. Such a period is unable to generate significant new density perturbations, but may be desirable in helping to alleviate some relic abundance problems not solved by the usual period of high-energy inflation. The second possibility is a period of matter-domination brought on by a long-lived massive particle, whose eventual decays restore radiation domination before nucleosynthesis. For definiteness, we shall take the long-lived particles to be the moduli fields of superstring theory, though the results apply for any non-relativistic decaying particle. ", "conclusions": "Although black hole constraints are an established part of modern cosmology, they are sensitive to the entire cosmological evolution. In the standard cosmology, a power-law spectrum is constrained to $n < 1.25$, presently the strongest observational constraint on $n$ from any source. Alternative cosmological histories can weaken this to $n < 1.30$, and worst-case non-gaussianity \\cite{BP} can weaken this by another 0.05 or so, though hybrid models giving constant $n$ give gaussian perturbations. Finally, we note that while the impact of the cosmological history on the density perturbation constraint is quite modest due to the exponential dependence of the formation rate, the change can be much more significant for other formation mechanisms, such as cosmic strings where the black hole formation rate is a power-law of the mass per unit length $G\\mu$. After all, the permitted initial mass density of black holes does increase by many orders of magnitude in these alternative cosmological models." }, "9804/astro-ph9804245_arXiv.txt": { "abstract": "We describe a method for the extraction of spectra from high dispersion objective prism plates. Our method is a catalogue driven plate solution approach, making use of the Right Ascension and Declination coordinates for the target objects. In contrast to existing methods of photographic plate reduction, we digitize the entire plate and extract spectra off-line. This approach has the advantages that it can be applied to CCD objective prism images, and spectra can be re-extracted (or additional spectra extracted) without having to re-scan the plate. After a brief initial interactive period, the subsequent reduction procedure is completely automatic, resulting in fully-reduced, wavelength justified spectra. We also discuss a method of removing stellar continua using a combination of non-linear filtering algorithms. The method described is used to extract over 12,000 spectra from a set of 92 objective prism plates. These spectra are used in an associated project to develop automated spectral classifiers based on neural networks. ", "introduction": "The MK classification of stellar spectra (Morgan, Keenan \\& Kellman 1943\\nocite{morgan_43a}) has been an important tool in the workshop of stellar and galactic astronomers for more than a century. While improvements in astrophysical hardware have enabled the rapid observation of digital spectra, our ability to efficiently analyze and classify spectra has not kept pace. Traditional visual classification methods are clearly not feasible for large spectral surveys. In response to this, we have been working on a project to develop automated spectral classifiers (von Hippel et~al.\\ 1994; Bailer-Jones 1996; Bailer-Jones et~al.\\ 1997, 1998). These classifiers, which are based on supervised artificial neural networks, can rapidly classify large numbers of digital spectra. The development of these classification techniques has required a large, representative set of previously classified spectra. The most suitable data has been the spectra from the Michigan Spectral Survey (Houk 1994)\\nocite{houk_94a} and the accompanying MK spectral type and luminosity class classifications listed in the {\\it Michigan Henry Draper} (MHD) catalogue (Houk \\& Cowley 1975; Houk 1978, 1982; Houk \\& Smith-Moore 1988). \\nocite{houk_75a}\\nocite{houk_78a}\\nocite{houk_82a}\\nocite{houk_88a} This paper describes the data reduction techniques we developed to extract and process these spectra. ", "conclusions": "\\begin{figure} \\centerline{ \\psfig{figure=fig9.eps,width=0.5\\textwidth,angle=0} } \\caption{Distribution of spectral types for each luminosity class. The dotted line represent giants (III), the dashed line subgiants (IV) and the solid line dwarfs (V).} \\label{dist_B} \\end{figure} This paper has described a method for extracting spectra from objective prism images. The method has been developed for the reduction of a set of photographic objective prism plates, but because the spectral extraction and processing takes place entirely in software using the complete digitized plate, it can equally well be applied to CCD objective prism images. The extraction process is driven by a set of catalogue Right Ascension and Declination positions, so a direct image of each field is not required. After an initial interactive period taking one or two minutes, the subsequent reduction is automatic, taking approximately one hour on a modest-sized SUN Sparc IPX to process a single plate (i.e.\\ extract about 150 spectra). The reduction method described in this paper has been used to extract a set of over 12,000 high-quality spectra. From this, a subset of over 5,000 normal spectra was selected which had reliable two-dimensional (spectral type and luminosity class) classifications listed in the MHD catalogue. The frequency distribution of the various stellar classes in this set is shown in Figure~\\ref{dist_B}. This data set is used in accompanying papers to produce automated systems for classifying and physically parametrizing stellar spectra (Bailer-Jones et~al.\\ 1997, 1998). In the interests of extending spectral classification to more distant stellar populations, spectra of stars fainter than B $\\sim 12$ are required. This could be achieved with a CCD objective prism survey. Although the technique described can only extract objects with known Right Ascension and Declination coordinates, the HST Guide Star Catalogue (e.g.\\ Lasker et al.\\ 1990)\\nocite{lasker_90a}, which lists 19 million objects brighter than 16$^{th}$ magnitude, could be used as a driver for extraction. However, Bailer-Jones (unpublished, 1996) has also modified the method to extract unwidened spectra from CCD objective prism images in the absence of any coordinates, using an algorithm to locate local flux peaks. The method can be applied to spectra at different spectral resolutions and wavelength coverages, provided a suitable line exists for the second plate solution." }, "9804/astro-ph9804135_arXiv.txt": { "abstract": "We have undertaken near-continuous monitoring of the Seyfert 1 galaxy NGC 7469 in the X-ray with \\xte\\ over a $\\sim 30$~d baseline. The source shows strong variability with a root-mean-square (rms) amplitude of $\\sim 16$~per cent, and peak--to--peak variations of a factor of order 2. Simultaneous data over this period were obtained in the ultraviolet (UV) using \\iue, making this the most intensive X-ray UV/X-ray variability campaign performed for any active galaxy. Comparison of the continuum light curves reveals very similar amplitudes of variability, but different variability characteristics, with the X--rays showing much more rapid variations. The data are not strongly correlated at zero lag. The largest absolute value of the correlation coefficient occurs for an anticorrelation between the two bands, with the X-ray variations leading the UV by $\\sim 4$~d. The largest positive correlation is for the ultraviolet to lead the X-rays by $\\sim 4$~d. Neither option appears to be compatible with any simple interband transfer function. The peak positive correlation at $\\sim 4$d occurs because the more prominent peaks in the UV light curve appear to lead those in the X-rays by this amount. However, the minima of the light curves are near-simultaneous. These observations provide new constraints on theoretical models of the central regions of active galactic nuclei. Models in which the observed UV emission is produced solely by re-radiation of absorber X-rays are ruled out by our data, as are those in which the X-rays are produced solely by Compton upscattering of the observed UV component by a constant distribution of particles. New or more complex models must be sought to explain the data. We require at least two variability mechanisms, which have no simple relationship. We briefly explore means by which these observations could be reconciled with theoretical models. ", "introduction": "\\label{Sec:Introduction} The origin of the continuum emission of Active Galactic Nuclei (AGN) -- which covers an extremely broad band -- is not well understood. In a number of high luminosity sources, it appears that this emission peaks in the ultraviolet (UV), the so-called ``Big Blue Bump'' (Shields 1978; Malkan \\& Sargent 1982). Strong, and apparently non-thermal X-ray flux is also a persistent property of AGN (e.g., Marshall \\etal\\ 1981). The X-ray emission covers a wide band from at least 0.1-100~keV, and can be described by a power-law form (Mushotzky, Done \\& Pounds 1993). The Big Blue Bump is often identified as the thermal output of an accretion disk (henceforth the accretion disk model e.g., Shakura \\& Sunyaev 1973). Heat is generated by viscous dissipation in the disk, which then radiates in the optical/UV regime for black hole masses typical of AGN (e.g., Sun \\& Malkan 1989). An alternative origin for the UV continuum emissions has been suggested by both observational and theoretical considerations. Guilbert \\& Rees (1988) postulated that the UV need not be internally generated in the accretion disk, but could arise via absorption and thermal re-emission (hereafter referred to as ``thermal reprocessing'') of X-rays in optically thick gas close to the central engine. The material - which could be but does not necessarily have to be the disk -- would imprint features on the X-ray spectra (e.g., Lightman \\& White 1988; George \\& Fabian 1991; Matt, Perola \\& Piro 1991). Such features have been found (e.g., Nandra \\& Pounds 1994) and suggest that approximately half of the incident X-rays are absorbed in the optically thick material. Spectroscopic observations of strong gravitational and Doppler effects in the iron K$\\alpha$ line profiles of Seyfert galaxies (e.g., Tanaka \\etal\\ 1995; Nandra \\etal\\ 1997) suggest that this material lies extremely close to the central black hole and is probably in the form of a disk (Fabian \\etal\\ 1995). The bulk of the continuum photons absorbed in the gas should then be re-emitted at the characteristic thermal temperature of the material. For dense gas close to the central engine, and particularly for ``standard'' accretion disks, this should be in the optical/UV. Most models for the X--ray continuum of AGN are based on the idea that lower-energy photons are Compton scattered by a population of hot electrons and/or pairs (which we refer to as ``upscattering'' models e.g., Sunyaev \\& Titarchuk 1980; Svensson 1983; Guilbert, Fabian \\& Rees 1983). The seed photons are often assumed to be those in the blue bump. Specific models differ primarily in their assumptions about, e.g., the geometry of the system (e.g. Haardt \\& Maraschi 1991, 1993; Haardt, Maraschi \\& Ghisellini 1994; Stern \\etal\\ 1995), the question of whether the electron population has a thermal or non-thermal distribution, the importance of pairs (e.g., Zdziarski \\etal\\ 1990, 1994). These models have been successful in explaining various observations. The goal underlying the exploration of the models is the discovery of the process responsible for the generation of the copious energy output of AGN. While the case for accreting supermassive black holes is becoming compelling, the method by which the rest-mass energy of the material is converted into radiative energy is still highly uncertain. Some specific questions which remain about the emission mechanisms include: \\begin{enumerate} \\item{How important is viscous dissipation in the generation of the UV?} \\item{What proportion of the UV arises via thermal reprocessing of X--rays?} \\item{What is the seed population for upscattering into the X--rays?} \\item{What mechanism accelerates the particles which up-scatter these seed photons?} \\end{enumerate} A powerful way of investigating these questions is by variability campaigns. These have already reaped rich rewards in the study of AGN emission lines via ``reverberation mapping'' (e.g. Peterson 1993, Netzer \\& Peterson 1997, and references therein). These emission line campaigns, however, also had strong implications for the generation of the continua, which we shall discuss below. The models discussed above all imply strong connections between the continuum emission in different bands. For example, the accretion disk emission could cover an extremely broad band, depending on the temperature profile of the disk. The thermal reprocessing model predicts that the X-rays should be generating UV emission. The upscattering model suggests the converse. By observing the variability in these bands, therefore, we can make inferences as to which of the various processes is in operation and to what degree. In particular, simultaneous X-ray/UV data should be the most revealing. A number of AGN have been monitored simultaneously at optical/UV and X-ray energies. Leaving aside blazars, the best-studied sources are NGC~4151, NGC~5548 and NGC 4051. In the first two objects, there is evidence for a correlation between the two bands. The best-sampled (and therefore most reliable) case is NGC~4151, in which the 1455~\\AA\\ and 2--10~keV flux appears to correlate well on all time scales from hours to a year (Perola \\etal\\ 1986; Edelson \\etal\\ 1996). In NGC~5548, the flux in the two bands is also well correlated on time scales from days to 1 year (Clavel \\etal\\ 1992). In both sources however, the correlation appears to break down during one very large UV outburst. NGC~4051 shows different behavior, in that the X-ray emission showed large-amplitude (factor $\\sim 2$) variability, while the optical emission remained steady to within a few per cent when observed over a $\\sim 2$~d baseline (Done \\etal\\ 1990). For completeness, we also mention the results obtained for other non-blazar AGN, though their significance is marginal due to the small number of simultaneous observations and/or the short duration of the campaigns. In Fairall~9, the slow decline of the 2--10~keV flux mimics the secular fading of the UV and optical continuum from 1978 to 1985 (Morini \\etal\\ 1986). The UV-optical versus X-ray flux correlation seems to hold in NGC~4593 (Santos-Ll\\'{e}o \\etal\\ 1995), whereas in MCG-8-11-11 (Treves \\etal\\ 1990), 3C120 (Maraschi \\etal\\ 1991) and 3C~273 (Courvoisier \\etal\\ 1990) the two wavebands appear to be independent of each other. These previous attempts at determining the relationship between the components have obviously left some ambiguity. This is perhaps not surprising as generally the sampling of the light curves has been rather poor. In order to provide an improved dataset, a campaign of near-continuous \\xte\\ and \\iue\\ monitoring of NGC 7469 was undertaken over a $\\sim 1$~month baseline. The results of the campaign in terms of the relationship of the X-ray and UV variability are the subject of this paper. We have effectively divided the paper into two halves. Sections 2-4 discuss the observational results exclusively, which are then summarized in Section 5. Section 6 then investigates the implications of the observational results within the framework of the models discussed above and suggest possible ways of reconciling the data with models. ", "conclusions": "We have investigated the relationship between the X-ray and UV emission in NGC 7469 on time scales of hours-weeks. The poor correlation between the X-ray and UV light curves at zero lag may be considered a surprising result because, as mentioned in \\S1, some previous experiments suggested a good correlation, and little if any time lag between the variations in the two bands (e.g. NGC 5548, NGC 4151). No other AGN, however, has been monitored as intensively and for such a long duration. Variability information has important implications for the physical mechanisms responsible for the production of the X-ray and UV emission in AGN. If the flux changes in two bands are correlated, this suggests some causal link between them. A time lag between the bands then shows which component drives the other. If X-ray variations lead those in the UV, this is strongly suggestive of thermal reprocessing. If the opposite is observed, this strongly favors upscattering. With a simple transfer function this delay should be similar for all ``events'' in the light curve. No such simple behavior was observed during our campaign and the interpretation is less straightforward. Our data require modifications to the simplest ideas about the emission mechanisms in the UV and X-ray. \\subsection{Model ingredients} It is widely accepted that the radiative energy from AGN originates as the rest-mass energy of accreting material. The conversion process must include a mechanism to accelerate the particles which produce the X-rays as they are the highest-energy photons to carry an appreciable fraction of the luminosity. In principle, the remainder of the AGN spectrum could be produced by thermal reprocessing following absorption of some fraction of this X-ray continuum. Viscous dissipation in the accretion flow, however, could dominate the observed UV emission. In many scenarios there is also a radiative connection between the UV and X-ray emission regions: X-rays can be produced by upscattering of UV seed photons and UV emission can be produced by re-radiation of absorbed X-rays. In these circumstances, a substantial number of factors can affect the observed variability: \\begin{enumerate} \\item Changes in the physical properties (e.g. optical depth, temperature, geometry) of the particle distribution which produces the X-rays \\item Instabilities in the accretion flow \\item The geometry and size of the X-ray and UV emission regions \\item Anisotropy of the radiation fields, which might include relativistic effects close to the black hole \\item The temperature distribution of the absorbing medium \\item The importance of feedback, in which variations in each band affect the other \\item Changes in occulting/absorbing media \\end{enumerate} \\subsection{Implications of the NGC 7469 data} Our data have a number of implications for the processes which produce the UV--to--X-ray emissions of NGC 7469 and the observed variability, which we now discuss. First we consider our observations in the context of historical data and the spectral energy distribution of NGC 7469. \\subsubsection{The broad-band perspective} The mean X-ray and UV fluxes observed for NGC 7469 during our campaign are very much typical of the respective historical means for this source. The mean flux in the 2--10 keV band, based on historical observations over the period 1979-1993, is $\\sim 3 \\times 10^{-11}$\\,erg cm$^{-2}$ s$^{-1}$. Furthermore, the range of historical variability is very similar to that observed during our campaign. This suggests that we sampled a large fraction of source variability in NGC 7469 during our one-month campaign, although we observe no obvious flattening of the PDS (Fig.~\\ref{fig:pds}) at the lowest frequencies. Chapman, Geller and Huchra (1985) derived a mean value of $4.6 \\times 10^{-14}$\\,erg~cm$^{-2}$~s$^{-1}$~\\AA$^{-1}$ for the 1430--1460~\\AA\\ continuum flux of NGC~7469 from 10 \\iue\\ observations in 1979--1982. Similarly, Edelson, Pike and Krolik (1990), reported a mean value of $4.8 \\times 10^{-14}$\\,erg~cm$^{-2}$~s$^{-1}$~\\AA$^{-1}$ for the continuum flux at 1450~\\AA\\ (rest wavelength) from 16 \\iue\\ observations in 1979--1985. With a mean UV flux at 1485~\\AA\\ (observed wavelength) of $4.0 \\times 10^{-14}$\\,erg~cm$^{-2}$~s$^{-1}$~\\AA$^{-1}$ (W97), NGC~7469 was thus neither particularly faint nor exceptionally bright during our campaign. The optical--to--X-ray spectral index of NGC~7469, $\\alpha_{ox}$~=~1.22, is not significantly different from that of, e.g., NGC~5548 ($\\alpha_{ox}$~=~1.25), or the mean index for Seyfert galaxies (Kriss, Canizares and Ricker 1980). The spectral energy distribution of NGC 7469 is shown in Fig.~\\ref{fig:sed}. The only real peculiarity of NGC~7469 is the presence of a circumnuclear starburst ring within $1.\\!''5$ of its nucleus. Genzel \\etal\\ (1995) estimate that the starburst accounts for two-thirds of the source bolometric luminosity and it may dominate the IR emission. However, it only contributes 4 percent to the observed X-ray flux (Perez-Olea and Colina 1996). The starburst should be invariant on the time scales sampled here and have no effect on the X--ray/UV variability. \\subsubsection{The X--ray continuum} The presence of rapid variations in the X-rays which are not seen in the UV implies that the particle distribution responsible for the X-rays is variable. The X-ray emission mechanism is highly uncertain, but as mentioned in the introduction, many models have concentrated on Compton upscattering of seed UV photons by a population of hot electrons and/or pairs (e.g. Haardt \\& Maraschi 1991, 1993). Our observations show that if the observed UV photons are the seed population, then the rapid variations of the X-rays do not arise from variations in the seed. Either the optical depth, temperature or geometry of the upscattering region must be changing. In the latter case, changes in the distribution of active regions in the X-ray source, or kinematic effects can produce variability (e.g., Abramowicz \\etal\\ 1991; Haardt, Maraschi \\& Ghisellini 1997). Longer-term variations are also observed in the X-ray flux and these could also arise from processes such as those just mentioned. The fact that the power spectrum shows no obvious features or break is supportive of this interpretation. The fluctuations have a similar amplitude to those in the UV and this suggests a connection between the bands. This is intriguing. One possibility is that the long time-scale variability in the X--rays is due to changes in the UV seed population. In the simplest such interpretation - where the upscattering region is point-like and lies in the line of sight to a point-like seed source we expect a 1:1 correlation between the two bands with no lag. This is ruled out by our data, although we note that the minima appear to be very close in time. In more complex geometries, there may be time lags which will be in the sense that the X-ray variations follow those in the UV. The delay of $\\sim 4$~d between the peaks is superficial evidence in favor of upscattering. In that model, however, we expect the lag to be very short, being dependent primarily on the light travel time between the regions, modified by geometrical factors. A lag of $\\sim 4$~d, the only plausible conclusion here, seems rather long to be associated with these processes. In addition, we are unable to envisage a purely geometrical modification of a linear process which accounts for {\\it both} the relationship between the maxima and that of the minima. \\subsubsection{The UV continuum} We now consider the alternative that the X-ray emission drives the UV in the thermal reprocessing scenario. The luminosities in the X-ray and UV bands are similar (Fig.~\\ref{fig:sed}) and thermal reprocessing can therefore be energetically important. As stated above, the 2-10 keV observed flux of NGC 7469 at our epoch was $3.4\\times 10^{-11}$\\,erg cm$^{-2}$ s$^{-1}$. However, the X-ray emission of NGC 7469 covers a far wider band than this, with significant emission being observed down to $\\sim 0.1$~keV with ROSAT (Brandt \\etal\\ 1993) and most likely up to at least 100~keV as seen by OSSE (Zdziarski \\etal\\ 1995; Gondek \\etal\\ 1996). Estimates of the underlying photon index of the continuum are in the range $1.9-2.0$ after accounting for the effects of Compton reflection (Piro \\etal\\ 1990; Nandra \\& Pounds 1994). We estimate the mean X--ray luminosity of NGC 7469 at our epoch to be $1.8\\times 10^{44}$\\,erg s$^{-1}$ in the 0.1-100~keV band. In the canonical thermal--reprocessing scenario about half of this luminosity should be absorbed in the accretion disk or other material. After estimating that fraction which is Compton scattered rather than absorbed (George \\& Fabian 1991) we conclude that the presence of the iron emission line and reflection hump in NGC 7469 indicate that a luminosity of $\\sim8\\times 10^{43}$\\,erg s$^{-1}$ of the X-ray emission of NGC 7469 is reprocessed and re-emerges as thermal emission. Let us now suppose that all of this luminosity emerges in a single black body (which represents the narrowest physically-realistic spectrum) peaking close to 1315\\AA\\ ($kT \\sim 2$~eV). Such a blackbody is almost sufficient to account for the continuum at 1315\\AA\\ (Figure~\\ref{fig:sed}). Therefore, it is energetically possible that reprocessed X-rays produce some of the observed UV continuum and its variations. As in the case of upscattering, the simplest thermal reprocessing models predict a strong positive correlation between the bands, with any time lags being in the sense that the X-ray variations lead those in the UV. No such lag is observed. We do find that the strongest {\\it anti}--correlation of the datasets occurs for the X-rays leading the UV, but the interpretation of such a result is far from obvious and we do not comment on it further. Even if aliasing has caused us to ``miss'' a positive correlation with a long ($\\sim 14$~d) lag between the UV and X--rays, any lag longer than a day or so is very difficult to explain. Any single transfer function which related the two bands would smooth the light curve of the responding band and reduce its amplitude. We observe, however, that the amplitudes of variability on long time scales are very similar. We therefore reject such a possibility. In this case it is even more difficult to envisage a complex geometry which can reproduce the light curves. It therefore seems highly unlikely that any substantial proportion of the 1315\\AA\\ continuum of NGC 7469 arises from thermal reprocessing unless, for example, there is substantial anisotropy of the X-ray emission. W97 found that the variations at longer UV wavelengths followed those at shorter wavelengths, but with a time lag of a fraction of a day. Collier \\etal\\ (1998) have demonstrated that this trend continues into the optical, and Peterson \\etal\\ (1998) find this trend to be significant at no less than the 97~per cent confidence level. This, together with the rapidity of the variations in NGC 7469, is most easily explicable in terms of the thermal reprocessing hypothesis. However, our data essentially rule out models in which all the observed optical/UV flux is re-radiated X-ray emission. The optical/UV variability therefore requires either intra-band reprocessing, which is difficult from an energetics standpoint, or some other model. Should we therefore conclude that the UV/optical continuum in NGC 7469 arises from direct emission by the accretion disk? Our data offer no direct constraints on accretion disk models, as no explicit relationship between the X--ray and UV emissions is predicted by those models. Nonetheless, the fact that thermal reprocessing is strongly disfavored by our data has profound implications for the disk models. The rapid and wavelength-coherent variations in the optical and UV flux of AGN is difficult to reconcile with a standard $\\alpha$-disk (e.g. Krolik \\etal\\ 1991; Molendi, Maraschi \\& Stella 1992). Prior to our observations, it was conceivable that thermal reprocessing was responsible for these rapid variations. There is now a clear need for a revision of accretion disk theory to account for these wavelength-independent variations without resorting to reprocessing. \\subsubsection{The extreme ultraviolet (EUV) continuum and UV emission lines} Given the presence of a typical iron K$\\alpha$ line and reflection hump in this source (Piro \\etal\\ 1990; Nandra \\& Pounds 1994) we are left with the question of where the putative reprocessed X-ray flux is emitted. One possibility is that the thermal reprocessing occurs in a molecular torus (Ghisellini, Haardt \\& Matt 1994; Krolik, Madau \\& Zycki 1994), in which case it might emerge in the infrared. Such an hypothesis would predict a narrow iron K$\\alpha$ line in the X-ray spectrum, whereas in many Seyfert galaxies these lines are extremely broad. The case of NGC 7469 is unclear, with Guainazzi \\etal\\ (1994) finding no evidence for a broad component and Nandra \\etal\\ (1997) finding marginal evidence. A conclusive determination requires a longer exposure with \\asca, but it seems highly likely that the iron K$\\alpha$ line and Compton hump in Seyfert 1 galaxies in general are produced extremely close to the central black hole (e.g., Nandra \\etal\\ 1997). We would therefore expect the reprocessed emission to emerge at a higher energy. As shown above, however, the thermally reprocessed X-rays only make a strong contribution to the observed optical and UV wavebands if the emission is strongly peaked at those wavelengths. It seems more likely that the emission covers a range of temperatures, in which case the reprocessed flux would be difficult to detect when spread over a wide band. Alternatively, it could peak in the (unobserved) EUV band. Figure~\\ref{fig:sed} shows that this can indeed be the case. A blackbody of luminosity $8\\times 10^{43}$\\,erg s$^{-1}$ contributes less than 5~per cent of the flux at 1315\\,\\AA\\ as long as $kT>12$~eV. Intensity variations in such a component would be undetectable with \\iue. Similarly, the \\asca\\ spectrum constrains $kT<60$~eV. If the spectral form is broader than a single blackbody, the range of allowed temperatures is correspondingly wider. Interestingly, Brandt \\etal\\ (1993) reported evidence for a soft excess in the \\ros/PSPC data which can be modeled as a blackbody of $kT\\sim 110$~eV and a luminosity of $10^{43}$\\,erg s$^{-1}$. This can be identified with the high energy tail of such a broad, reprocessed component. The major UV emission lines are excited by unobservable EUV photons. An extrapolation of the X-ray spectrum observed by \\asca\\ (George \\etal\\ 1998) and of the UV spectrum into that band indicate roughly equal contributions at energies at which the lines are excited. With the two components being poorly correlated at zero lag, it is therefore difficult to determine which will be the dominant EUV component at any given time. We have suggested above that there may even be a third contributor to the EUV, the reprocessed X-rays. In other words, the shape of the ionizing continuum changes with time. This effect could account for certain difficulties which have been encountered in explaining the emission line responses to the observed UV continuum in reverberation mapping experiments. Our observations suggest that the unseen EUV continuum is not directly related to the observed \\iue\\ flux, which therefore cannot be assumed to be a perfect representation of the continuum driving the line emission. It is also interesting to note that the emission-line light curves show long term trends which are not apparent in the continuum bands. This is most clearly demonstrated by Fig.~\\ref{fig:renorm_lc}, which shows the X-ray and UV continuum light curves, together with those of the Ly$\\alpha$ and C{\\sc iv} emission lines. These have all been renormalized to the $F_{\\rm var}$ value and thus the y-axes crudely represent the number of standard deviations from the mean. Both emission lines clearly show a long-term reduction in their flux which is not seen in either continuum. \\subsection{Steps towards a new model} In the light of the above, it is clear that new or more complex models must be sought to explain the data which have been obtained thus far, and particularly those described in this paper. Here we suggest some ways in which our new data might be reconciled with the existing paradigm by modification or extension. We emphasize that such a discussion is incomplete and {\\it ad-hoc}. As we have stated above, it seems most likely that the X--ray flux which is absorbed when the iron K$\\alpha$ lines is being generated emerges in a relatively weak, broad component, that may peak the EUV/soft X-ray band. The emission in this band may well provide the crucial connection between the higher and lower-energy components. A reasonable interpretation of the longer-timescale variability observed in our light curve is that the UV emission leads that in the X-rays, but with a variable lag. This suggests the dominant source of variations is in the seed population of an upscattering model. We do, however, bear in mind the caveat that the particle distribution of the upscattering medium must also be variable, to produce the most rapid variations. To explain the ``variable'' time lag, we suggest that there are multiple ``seed'' populations, which dominate at different times. In particular we suggest that the main source of 1315\\AA\\ photons is located at a distance of $\\sim 4$~lt d from the X--ray source and they are the dominant seed population when the source is in a high flux state, thus introducing a 4d ``lag'' between the X-rays and UV. When the 1315\\AA\\ flux is observed to decline, however, this allows other emitting regions to dominate the seed distribution. In particular we suggest at these times that EUV/soft X-ray photons are the dominant seed population. They arise from closer in and are therefore observed to have little or no lag with the X-rays. As might be apparent from the above discussion, the primary X--rays, reprocessed X-rays and primary UV might well exist in a rather fine balance in the typical AGN. Future observational data on other objects of similar quality to that presented here, and preferably including the far-UV and soft X-ray bands, will be necessary for further progress and to establish the generality of the phenomena explored here." }, "9804/astro-ph9804303_arXiv.txt": { "abstract": "We report a tentative detection with the IRAM 30m telescope of the LiH molecule in absorption in front of the lensed quasar B0218+357. We have searched for the $J = 0 \\rightarrow 1$ rotational line of lithium hydride at 444 GHz (redshifted to 263 GHz). The line, if detected, is optically thin, very narrow, and corresponds to a column density of N(LiH) = 1.6 10$^{12}$ cm$^{-2}$ for an assumed excitation temperature of 15 K, or a relative abundance LiH/H$_2 \\sim$ 3 10$^{-12}$. We discuss the implications of this result. ", "introduction": "Primordial molecules are thought to play a fundamental role in the early Universe, when stellar nucleosynthesis has not yet enriched the interstellar medium. After the decoupling of matter and radiation, the molecular radiative processes, and the formation of H$_2$, HD and LiH contribute significantly to the thermal evolution of the medium (e.g. Puy et al 1993, Haiman, Rees \\& Loeb 1996). Even at the present time, it would be essential to detect such primordial molecules, to trace H$_2$ in the low-metallicity regions (e.g. Pfenniger \\& Combes 1994, Combes \\& Pfenniger 1997). Unfortunately, the first transition of HD is at very high frequency (2.7 THz), and the first LiH line, although only at 444 GHz, is not accessible from the ground at $z=0$ due to H$_2$O atmospheric absorption. This has to wait the launching of a submillimeter satellite. Although the Li abundance is low (10$^{-10}$-10$^{-9}$), the observation of the LiH molecule in the cold interstellar medium looks promising, because it has a large dipole moment, $\\mu = 5.9$ Debye (Lawrence et al.~1963), and the first rotational level is at $\\approx 21\\,\\rm K$ above the ground level, the corresponding wavelength is $0.67\\,\\rm mm$ (Pearson \\& Gordy 1969; Rothstein 1969). The line frequencies in the submillimeter and far-infrared domain have been recently determined with high precision in the laboratory (Plummer et al 1984, Bellini et al 1994). Because of the great astrophysical interest of this molecule (e.g. Puy et al 1993), an attempt has been made to detect LiH at very high redshifts ($z \\sim 200$) with the IRAM 30m telescope (de Bernardis et al 1993). It has been proposed that the LiH molecules could smooth the primary CBR (Cosmic Background Radiation) anisotropies, due to resonant scattering, or create secondary anisotropies, and they could be the best way to detect primordial clouds as they turn-around from expansion (Maoli et al 1996, but see also Stancil et al 1996, Bougleux \\& Galli 1997). There has recently been some controversy about the abundance of LiH. The computations of Lepp \\& Shull (1984) estimated the LiH/H$_2$ abundance ratio in primordial diffuse clouds to be as high as 10$^{-6.5}$. With H$_2$/H $\\sim$ 10$^{-6}$, the primordial LiH/H ratio is $\\sim$10$^{-12.5}$. More recently, Stancil et al. (1996) computed an LiH/H abundance of $< 10^{-15}$ in the postrecombination epoch, since quantum mechanical computations now predict the rate coefficient for LiH formation through radiative association to be 3 orders of magnitude smaller than previously thought from semi-classical methods (Dalgarno et al 1996). In very dense clouds, however, three-body association reactions must be taken into account, and a significant fraction of all lithium will turn into molecules. Complete conversion due to this process requires gas densities of the order $\\sim 10^9$\\,cm$^{-3}$, rarely found in the general ISM. However, taken other processes into account, such as dust grain formation, an upper limit to the LiH abundance is the complete conversion of all Li into molecular form, with LiH/H$_2$ $\\la 10^{-10}-10^{-9}$. With a LiH column density of 10$^{12}$\\,cm$^{-2}$, or N(H$_2$)$= 10^{22}$\\,cm$^{-2}$, the optical depth of the LiH line will reach $\\sim$1, in cold clouds of velocity dispersion of $2\\,\\rm km\\,s^{-1}$. The line should then be easily detectable in dense dark clouds in the present interstellar medium (like Orion where the column density reaches 10$^{23}$-10$^{24}$ cm$^{-2}$). This is a fundamental step to understand the LiH molecule formation, in order to interpret future results on primordial clouds, although the primordial abundance of Li could be increased by about a factor 10 in stellar nucleosynthesis (e.g. Reeves 1994). Once the Li abundance is known as a function of redshift, it could be possible to derive its true primordial abundance, a key factor to test Big Bang nucleosynthesis (either homogeneous or not). Up to now, due to atmospheric opacity, no astrophysical LiH line has been detected, and the abundance of LiH in the ISM is unknown. The atmosphere would allow to detect the isotopic molecule LiD (its fundamental rotational line is at 251 GHz), but it has not been seen because of the low D/H ratio, and the expected insufficient optical depth of LiH \\footnote{the LiD line at 251 GHz is not covered in the 247-263 GHz survey of Orion by Blake et al 1986, but was observed at the McDonald 5m-telescope, Texas, see Lovas 1992; we have ourselves checked with the SEST telescope that no line is detected towards Sagittarius-B2 at this frequency. The 3$\\sigma$ upper limit to the LiD abundance towards SgrB2 is $1 \\times 10^{11}$\\,cm$^{-2}$.}. Another method to avoid atmospheric absorption lines is to observe a remote object, for which the lines are redshifted into an atmospheric window. Here we report about the first absorption search for a LiH line at high redshift: the latter allows us to overcome the earth atmosphere opacity, and thanks to the absorption technique we benefit of an excellent spatial resolution, equal to the angular size of the B0218+357 quasar core, of the order of 1 milli-arcsec (Patnaik et al 1995). At the distance of the absorber (redshift $z=0.68466$, giving an angular size distance of 1089 Mpc, for $H_0$=75 km/s/Mpc and $q_0$ =0.5), this corresponds to 5pc. We expect a detectable LiH signal, since the H$_2$ column density is estimated to be N(H$_2$)$ = 5 \\times 10^{23}$ cm$^{-2}$. Menten \\& Reid (1996) derive an N(H$_2$) value ten times lower than this, using the H$_2$CO($2_{11}-2_{12}$) transition at 8.6\\,GHz. At this low frequency the structure and extent of the background continuum source may be quite larger than at 100--200 GHz and the source covering factor smaller. This means that their estimate of the column density is a lower limit. \\section { Observations } The observations were made with the IRAM 30m telescope at Pico Veleta near Granada, Spain. They were carried out in four observing runs, in December 1996, March, July and December 1997. Table 1 displays the observational parameters. We observed at 263 GHz with an SiS receiver tuned in single sideband (SSB). The SSB receiver temperature varied between 400 and 450K, the system temperature was 600-1400K depending on weather conditions, and the sideband rejection ratio was 10dB (the image frequency is at 271.5 GHz, in a region where the atmospheric opacity increases rapidly due to water vapour). We used a 512x1MHz filterbank and an autocorelator backend, with 0.3 km/s resolution. We present here only the 1MHz resolution spectra, smoothed to a 2.3 km/s channels, to improve the signal to noise. \\begin{table} \\begin{flushleft} \\caption[]{ Parameters for the tentative LiH line } \\begin{tabular}{lccl} \\hline \\multicolumn{1}{l}{J$_u$--J$_l$ } & \\multicolumn{1}{c}{1--0 } & \\multicolumn{1}{c}{} \\\\ $\\nu_{lab}$ GHz\t& 443.953 & \\\\ $\\nu_{obs}$ GHz\t& 263.527 & \\\\ Forward eff.\t & 0.86 &\t\\\\ Beam eff.\t & 0.32 & \\\\ T$_A^*$ & 7 mK\t& depth of absorption line \\\\ T$_{\\rm cont}$ & 15 mK & \\\\ FWHM \t & 3.2\tkm/s & \\\\ $\\sigma$ & 1.8 mK\t& noise rms with $\\Delta v$ 2.3 km/s \\\\ \\hline \\end{tabular} \\, \\\\ \\vskip 2truemm $\\alpha$(1950) = 02h 18m 04.1s \\\\ $\\delta$(1950) = 35$^\\circ$ 42\\amin \\, 32\\asec \\\\ \\end{flushleft} \\end{table} The observations were done using a nutating subreflector with a 1' beamthrow in azimuth. We calibrated the temperature scale every 10 minutes by a chopper wheel on an ambient temperature load, and on liquid nitrogen. Pointing was checked on broadband continuum sources, and was accurate to 3\\asec \\, rms. The frequency tuning and sideband rejection ratios were checked by observing molecular lines towards Orion, DR21 and IRC+10216. We integrated in total for 85 hours on the 263 GHz line, and obtained a noise rms level of 1.8 mK in the T$_A^*$ antenna temperature scale, with a velocity resolution of 2.3 km/s. The forward and beam efficiencies at the observed frequency are displayed in Table 1. The continuum level was estimated by observing in a rapid on--off mode using a special continuum backend. The switch frequency of the subreflector was increased from 0.5 Hz to 2 Hz. \\begin{figure} \\psfig{figure=Ad212_f1.ps,bbllx=3cm,bblly=65mm,bburx=11cm,bbury=195mm,width=8cm} \\caption[]{ Spectrum of LiH in its fundamental line (1--0) at 444 GHz, redshifted at 263 GHz, in absorption towards B0218+357, compared to the highly optically thick CO(2--1) line previously detected. The tentative LiH line is slightly shifted from the center by about 5km/s, but is still comprised within the CO(2--1) velocity range. Its width is compatible with what is expected from an optically thin line. Spectra have been normalised to the absorbed continuum level and the velocity resolution is 2.3 km/s } \\label{lih_f1} \\end{figure} ", "conclusions": "Figure \\ref{lih_f1} presents our LiH spectrum, compared to that of CO(2--1) previously detected with the IRAM 30m-telescope (Wiklind \\& Combes 1995, Combes \\& Wiklind 1995). There is only a tentative detection of LiH at $\\sim$ 3 $\\sigma$. The line is very narrow, but is compatible to what is expected from an optically thin line. The CO(2--1) is highly optically thick, with $\\tau \\sim$ 1500. This optical depth is determined from the detection of C$^{18}$O(2--1), which is moderately thick, and the non--detection of C$^{17}$O(2--1). The center of the tentative line is shifted by 5 km/s from the average center of other lines detected towards B0218+357. This shift cannot be attributed to uncertainties of the line frequency, since it has been measured in the laboratory (e.g. Bellini et al 1994), and the error is at most 0.24 km/s at 3$\\sigma$, once redshifted. But the scatter of the line centers is $\\sim$ 3 km/s, and the width of most of the lines is $\\sim$ 15 km/s (cf Wiklind \\& Combes 1998). The velocity shift is therefore insufficient to reject the line as real. Combining our own continuum data with that of lower frequencies (obtained from the NASA Extragalactic Database NED), we have previously found that the continuum spectra of B0218+357 can be fitted with a power law of slope --0.25 (Combes \\& Wiklind 1997). This would imply a continuum level of 15.5 mK at 263 GHz, which is in accord with the measured level. Since only 70\\% of the continuum is covered by molecular gas, the continuum level to be used for our LiH observations amounts to 11 mK. \\smallskip We can write the general formula, concerning the total column density of the LiH molecule, observed in absorption between the levels $l \\rightarrow u$ with an optical depth $\\tau$ at the center of the observed line of width $\\Delta v$ at half-power: $$ N_{LiH} = {{8\\pi}\\over{c^3}} f(T_x) {{\\nu^3 \\tau \\Delta v} \\over {g_u A_u} } $$ where $\\nu$ is the frequency of the transition, $g_u$ the statistical weight of the upper level ($= 2 J_u+1$), $A_u$ the Einstein coefficient of the transition, $T_x$ the excitation temperature, and $$ f(T_x) = {{Q(T_x) exp(E_l/kT_x)} \\over { 1 - exp(-h\\nu/kT_x)}} $$ where $Q(T_x)$ is the partition function. For the sake of simplicity, we adopt the hypothesis of restricted Thermodynamical Equilibrium conditions, i.e. that the excitation temperature is the same for all the LiH ladder. Since the line is not optically thick, but the optical thickness reaches $\\tau$ = 1.3 at the center of the line, we have derived directly from the spectrum, through a Gaussian fit of the opacity, the integrated $\\tau \\Delta v$ = 3.64 km/s. From the formulae above, and assuming an excitation temperature of $T_x$ = 15 K (see Table 2 for variation of this quantity), we derive a total LiH column density of 1.6 10$^{12}$ cm$^{-2}$ towards B0218+357. Compared to our previously derived H$_2$ column density of 5 10$^{23}$ cm$^{-2}$, this gives a relative abundance of LiH/H$_2$ $\\sim$ 3 10$^{-12}$. Note that there is a possible systematic uncertainty associated with this measure, due to the velocity difference between the maximum opacity of the CO, HCO$^+$ and other lines with that of LiH. \\begin{table} \\begin{flushleft} \\caption[]{Derived LiH column density } \\begin{tabular}{lccccc} \\hline & & & & & \\\\ \\multicolumn{1}{c}{$T_x$ } & \\multicolumn{1}{c}{(K)} & \\multicolumn{1}{c}{5 } & \\multicolumn{1}{c}{10 } & \\multicolumn{1}{c}{15 } & \\multicolumn{1}{c}{20 } \\\\ & & & & & \\\\ \\hline & & & & & \\\\ N(LiH) & (10$^{12}$ cm$^{-2}$) & 0.4 & 0.9 & 1.6 & 2.4 \\\\ & & & & & \\\\ LiH/H$_2$ & (10$^{-12}$) & 0.8 & 1.8 & 3.2 & 5 \\\\ & & & & & \\\\ \\hline \\end{tabular} \\end{flushleft} \\end{table} \\smallskip To interpret this result, comparison should be made with the atomic species. First, it is likely that the molecular cloud on the line of sight is dense and dark, and all the hydrogen is molecular, f(H$_2$) = 0.5. The Li abundance (main isotope $^7$Li) at $z=0.68466$ (i.e 5-10 Gyr ago) can be estimated at Li/H $\\sim 10^{-9}$, since its abundance in the ISM increases with time. The primordial Li abundance must be similar to that in metal deficient unevolved Population II stars, Li/H = 1-2 10$^{-10}$ (Spite \\& Spite 1982), but Li could be depleted at the stellar surface by internal mixing. In meteorites and unevolved, unmixed Pop I stars, Li/H $\\sim 10^{-9}$, representative of the Li abundance some 4 Gyr ago. The present abundance in the ISM is estimated around 3 10$^{-9}$ (Lemoine et al 1993). We therefore deduce LiH/Li $\\sim$ 1.5 10$^{-3}$. The uncertainty associated with the derived abundances are large, but the low LiH/Li ratio seems to exclude complete transformation of Li into LiH, as would be expected in very dense clouds (e.g. Stancil et al 1996, although the Li chemistry is not yet completely understood in dark clouds). However, it is likely that the cloud is clumpy, and in some of the more diffuse parts, LiH is photodissociated (e.g. Kirby \\& Dalgarno 1978). Also, some regions of the cloud could have a higher excitation temperature, in which case our computation under-estimates the LiH abundance (although the absorption technique selects preferentially cold gas, and the black-body temperature at the redshift of the absorbing molecules is $T_{bg}$ = 4.6 K). The present observations suggest that the detection of LiH in emission towards dense clouds in the Milky Way should be easy with a submillimeter satellite, provided that the spatial resolution is enough to avoid dilution of the dense clumps. It is also interesting to observe the rarer molecule $^6$LiH, which in some clouds might be of same order of abundance as the main isotopic species. Through optical absorption lines Lemoine et al (1995) find towards two velocity components in $\\zeta$-Oph, $^7$Li/$^6$Li = 8.6 and 1.4. Since $^6$Li is formed only in negligible amounts in the Big Bang, this ratio indicates that cosmic ray spallation has increased significantly the Li abundances. \\vspace{0.25cm}" }, "9804/astro-ph9804186_arXiv.txt": { "abstract": "We analysed 13 archival \\R\\ PSPC and HRI observations which included the position of a newly discovered 59\\,s X--ray pulsar in the Small Magellanic Cloud, 1SAX J0054.9--7226 = \\src. The source was detected three times between 1991 and 1996 at a luminosity level ranging from $\\sim$8$\\times$10$^{34}$ - 4$\\times$ 10$^{35}$ erg s$^{-1}$ (0.1--2.4 keV). Highly significant pulsations at 59.072\\,s were detected during the 1991 October 8--9 observation. The \\R\\ period, together with those measured by \\RXTE\\ and \\BSAX\\ yields a period derivative of \\.P= -- 0.016 s yr$^{-1}$. A much more accurate source position (10$^{\\prime\\prime}$ uncertainty) was obtained through the \\R\\ HRI detection on 1996 April restricting to three m$_ V$ $>$ 15.5 stars the likely counterpart of 1SAX J0054.9--7226 = \\src. ", "introduction": "On 1998 January 20 during a \\RXTE\\ observation in the direction of the Small Magellanic Cloud (SMC), a previously unknown X--ray source, namely XTE J0055--724, was detected at a flux level (2--10 keV) of $\\sim$ 6.0 $\\times$ 10$^{-11}$ erg s$^{-1}$ cm$^{-2}$. The source showed pulsations at a period of $\\sim$59\\,s (Marshall \\& Lochner, 1998a). A previous \\RXTE\\ observation of the same field performed on 1998 January 12 failed to detect the source. In response to these findings, simultaneous \\BSAX\\ and \\RXTE\\ observations of a region including the \\RXTE\\ error circle ($\\sim$10$^{\\prime}$ radius) of XTE J0055--724, were carried out on 1998 Jannuary 28. The results of these observations are reported elsewhere (Santangelo \\etal 1998a; Marshall \\etal 1998b). Thanks to the spatial capabilities of the imaging X--ray concentrators on board \\BSAX, an improved position ($\\sim$40$^{''}$ radius) was obtained for the pulsating source, named 1SAX J0054.9--7226 (Santangelo \\etal 1998a,b). The new \\BSAX\\ error circle contains only the previously classified \\R\\ and \\E\\ X--ray sources 1WGA J0054.9--7226 and \\src, which are likely the same source. In the following we adopt the earliest source name, i.e. \\E's. \\src\\ is a variable X--ray source in the SMC, which was already considered a candidate High Mass X--ray Binary by Wang \\& Wu (1992; source \\#35), Bruhweiler et al. (1987; source \\# 9) and by White \\etal (1994; in the WGACAT), based on its high spectral hardness. We report in this letter on the results from the analysis of the Position Sensitive Proportional Counter (PSPC) and High Resolution Imager (HRI) observations from the \\R\\ public archive. ", "conclusions": "\\src\\ was detected three times between 1991 and 1996 in the \\R\\ archival data. Highly significant pulsations, at a period of 59.072\\,s were detected on 1991 October 8--9. These findings, together with the \\BSAX\\ results, yield a mean period derivative of $\\sim$ -- 0.016\\, s yr$^{-1}$ between 1991 and 1998. \\begin{figure}[tbh] \\centerline{\\psfig{figure=59s_dss.ps,width=8.cm,height=8.cm}} \\caption{ESO plate including the position of \\src. The X--ray error circles obtained from different instruments and satellites are shown} \\end{figure} In one case a spectral analysis could be performed. The spectrum was found to be consistent with a relatively flat low absorbed power--law model that is typical of accreting X--ray pulsars in this energy range. The 0.1--2.4 keV luminosity of \\src\\ as observed with \\R\\ ranges between $\\sim$4.2$\\times$10$^{35}$ erg s$^{-1}$ (1991 October 8--9) and $\\sim$8.5$\\times$10$^{34}$ erg s$^{-1}$ (1996 April 26 -- June 10). Moreover \\RXTE\\ detected \\src\\ at a luminosity level of $\\sim$3$\\times$10$^{37}$ erg s$^{-1}$ in the 2--10 keV energy band. Extrapolating to the \\R\\ energy range the luminosity measured by \\RXTE\\ on 1998 January 20, a 0.1--2.4 keV luminosity of $\\sim$2.5$\\times$10$^{36}$ erg s$^{-1}$ is derived, implying a pronounced long--term variability of \\src\\ (a factor of $>$30). This indicates that the source is probably a transient X--ray pulsar in a high--mass binary containing a Be star. A 10$^{\\prime\\prime}$ accurate position was obtained thanks to a \\R\\ HRI observation during which the source was detected (1996 April; 0.1--2.4 keV luminosity of $\\sim$8.5$\\times$10$^{34}$ erg s$^{-1}$). The \\R\\ HRI error circle of contains only three stars in the ESO plates with m$_V$ $>$ 15.5, the likely optical counterpart of \\src (see Fig.\\,3). Assuming a B--V = --0.2 and a distance modulus of 19 mag, these optical counterpart candidates are consistent with main sequence A9 -- B2 stars. We note that a similar spectral--type star (B1.5Ve; m$_V$ = 16) is the companion of the nearby X--ray source SMC X--2. Future optical follow--up observations of these candidates should determine the counterpart of \\src\\ and its probable Be star X--ray transient nature. The optical and/or infrared activity brightening of the counterpart will allow further X--ray triggers and studies." }, "9804/astro-ph9804009_arXiv.txt": { "abstract": "The interstellar cloud surrounding the solar system regulates the galactic environment of the Sun, and determines the boundary conditions of the heliosphere. Both the Sun and interstellar clouds move through space, so these boundary conditions change with time. Data and theoretical models now support densities in the cloud surrounding the solar system of n(H$^{\\circ}$)=0.22$\\pm$0.06 cm$^{-3}$, and n(e$^{-}$)$\\sim$0.1 cm$^{-3}$, with larger values allowed for n(H$^{\\circ}$) by radiative transfer considerations. Ulysses and Extreme Ultraviolet Explorer satellite He$^{\\circ}$ data yield a cloud temperature of {\\mbox 6,400 K}. Nearby interstellar gas appears to be structured and inhomogeneous. The interstellar gas in the Local Fluff cloud complex exhibits elemental abundance patterns in which refractory elements are enhanced over the depleted abundances found in cold disk gas. Within a few parsecs of the Sun, inconclusive evidence for factors of 2--5 variation in Mg$^{+}$ and Fe$^{+}$ gas phase abundances is found, providing evidence for variable grain destruction. In principle, photoionization calculations for the surrounding cloud can be compared with elemental abundances found in the pickup ion and anomalous cosmic ray populations to model cloud properties, including ionization, reference abundances, and radiation field. Observations of the hydrogen pile-up at the nose of the heliosphere are consistent with a barely subsonic motion of the heliosphere with respect to the surrounding interstellar cloud. Uncertainties on the velocity vector of the cloud that surrounds the solar system indicate that it is uncertain as to whether the Sun and $\\alpha$ Cen are or are not immersed in the same interstellar cloud. ", "introduction": "The physical conditions of the surrounding interstellar cloud establish the boundary conditions of the solar system and heliosphere. The abundances and ionization states of elements in the surrounding interstellar cloud determine the properties of the parent population of the anomalous cosmic ray and pickup ion components. In addition, the history of the interstellar environment of the heliosphere appears to be partially recorded by radionucleotides such as $^{10}$Be and $^{14}$C in geologic ice core records (\\cite{sonett,fr97}). Because the solar wind density decreases as $R^{-2}$ ($R$=distance to Sun), the solar wind and interstellar densities are equal at about 5 AU (the orbit of Jupiter), in the absence of substantial ``filtration'' \\footnote{``Filtration'' refers to the deflection of interstellar H$^{\\circ}$ around the heliopause due to the coupling between interstellar protons and H$^{\\circ}$ resulting from charge exchange}. Approximately 98\\% of the diffuse material in the heliosphere is interstellar gas (\\cite{gruntman}). Thus, the physical properties of the outer heliosphere are dominated by interstellar matter (ISM). Were the Sun to encounter a high density interstellar cloud, it is anticipated that the physical properties of the inner heliosphere would also be ISM-dominated. Zank and Frisch (1998) have shown that if the space density of the interstellar cloud which surrounds the solar system were increased to $\\sim$10 cm$^{-3}$, the properties of the inner heliosphere at the 1 AU position of the Earth would be dramatically altered. The accuracy with which the physical properties of the surrounding cloud can be derived from observations of stars within a few parsecs of the Sun (1 pc$\\sim$200,000 AU) depends on the homogeneity and physical parameters of nearby ISM. Observations of nearby stars gives sightlines which probe the ensemble of nearby clouds constituting the ``Local Fluff'' cloud complex. Conclusions based on observations of nearby stars, however, must be qualified by the absence of detailed data pertaining to the small scale structure of the local ISM (LISM). More distant cold diffuse interstellar gas is highly structured, replete with dense ($\\sim 10^{4}-10^{5}$ cm$^{-3}$), small (20--200 AU) inclusions occupying in some cases less than 1\\% of the cloud volume (\\cite{frail,falgarone,falpug,heiles}). Small scale structures are ubiquitous in interstellar gas, and individual velocity components exhibiting column densities as low as N(H$^{\\circ}$)$\\sim$3$\\times 10 ^{18}$ cm$^{-3}$ are found in cold clouds (\\cite{frail,heiles}). The presence of dense low column density wisps near the Sun is allowed by currently available data. The Sun has a peculiar motion with respect to the ``Local Standard of Rest'' (LSR\\footnote{The LSR is the velocity frame of reference in which the vector motions of a group of nearby comparison stars are minimized. Stars in the LSR corotate around the galactic center with a velocity of $\\sim$250 km s$^{-1}$}); the Sun moves through the LSR with a velocity V$\\sim$16.5 km s$^{-1}$ towards the apex direction l=53$^{\\circ}$, b=+25$^{\\circ}$ (\\cite{mihalas}). Uncertainties on the relative solar-LSR motion appear to be less than 3 km s$^{-1}$ and $\\pm$5$^{\\circ}$. This motion corresponds to $\\sim$17 pc per million years. Note that the solar path is tilted by $\\sim25^{\\circ}$ with respect to the galactic plane. The Sun oscillates about the galactic plane, crossing the plane every 33 Myrs, reaching a maximum distance from the plane of $\\sim$77 pc. The last galactic plane ``crossing'' was about 21 Myrs ago (\\cite{bash}). This amplitude of oscillation can be compared to scale heights on the order of $\\sim$50-80 pc for cold H$_{2}$ and CO, $\\sim$100 pc for cold H$^{\\circ}$ and infrared cirrus, $\\sim$250 pc for warm H$^{\\circ}$, and $\\sim$1 kpc for warm H$^{+}$ (the ``Reynolds Layer''). There are three time scales of interest in understanding the environmental history of the solar galactic milieu -- $\\sim 10^{6}$ years, $\\sim 10^{5}$ years, and $\\sim 10^{4}$ years. Prior to entering into the Local Fluff complex of interstellar clouds, the Sun traveled through a region of the galaxy between the Orion spiral arm and the spiral arm spur known as the Local Arm. On the order of a million years ago, the Sun was displaced $\\sim$17 pc in the anti-apex direction, which is towards the present day location of the junction of the borders of the constellations of Columba, Lepus and Canis Major. The motions of the Sun and surrounding interstellar cloud with respect to interstellar matter within 500 pc, projected onto the plane, are illustrated in Figure 1. Note that the velocity vectors of the Sun and interstellar cloud surrounding the solar system are nearly perpendicular in the LSR, implying that the surrounding cloud complex is sweeping past the Sun (see section \\ref{velocity}). When the morphology of the Local Fluff complex is considered, it is apparent that sometime during the past $\\sim$200,000 years the Sun appears to have emerged from a region of space with virtually no interstellar matter (densities n(H$^{\\circ})<0.0005$ cm$^{-3}$, n(e$^{-})<0.02$ cm$^{-3}$) and entered the Local Fluff complex of clouds (average densities n(H$^{\\circ}$)$\\sim$0.1 cm$^{-3}$) outflowing from the Scorpius-Centaurus Association of star-forming regions. One model for the morphology of the cloud surrounding the solar system predicts that sometime within the past 10,000 years, and possibly within the past 2,000 years, the Sun appears to have entered the interstellar cloud in which it is currently situated (\\cite{fr94}, Frisch 1997). The cloud surrounding the solar system will be called here the ``Local Interstellar Cloud'' (LIC\\footnote{This cloud surrounding the solar system is also referred to as the ``surrounding interstellar cloud'', or SIC, which unambiguously defines the cloud feeding interstellar matter into the solar system. For the sake of uniformity of notation, however, the term LIC is used here.}). \\begin{figure} \\begin{center} \\plotone{fig1.eps} \\end{center} \\vspace{0.5in} \\caption[]{{\\small The distribution of interstellar molecular clouds (traced by the CO 1-$>$0 115 GHz rotational transition) and diffuse gas (traced by E(B-V) color excess due to the reddening of starlight by interstellar dust) within 500 pc of the Sun are shown. The round circles are molecular clouds, and the shaded material is diffuse gas. The horizontal bar (lower left) illustrates a distance of 100 pc. Interstellar matter is shown projected onto the galactic plane, and the plot is labeled with galactic longitudes. The distribution of nearby interstellar matter is associated with the local galactic feature known as ``Gould's Belt'', which is tilted by about 15--20$^{\\circ}$ with respect to the galactic plane. ISM towards Orion is over 15$^{\\circ}$ below the plane, while the Scorpius-Centaurus material (longitudes 300$^{\\circ}$--0$^{\\circ}$) is about 15--20$^{\\circ}$ above the plane. Also illustrated are the space motions of the Sun and local interstellar gas, which are nearly perpendicular in the LSR velocity frame. The three asterisks are three subgroups of the Scorpius-Centaurus Association. The three-sided star is the Geminga Pulsar. The arc towards Orion represents the Orion's Cloak supernova remnant shell. The other arcs are illustrative of superbubble shells from star formation in the Scorpius-Centaurus Association subgroups. The smallest (i. e. greatest curvature) shell feature represents the Loop I supernova remnant.}} \\label{fig1} \\end{figure} ", "conclusions": "One new conclusion presented here is that in principle the {\\it in situ} pickup ion data can help resolve the outstanding question of whether the correct reference abundances for the LIC are given by solar versus B-star abundances. A second new result is that the uncertainties on the LIC velocity vector indicate that it is not yet clear whether the Sun and $\\alpha$ Cen are immersed in the same interstellar cloud. Based on the discussions in this paper, the best values for LIC properties are given by n(H$^{\\circ}$)=0.22$\\pm$0.06 cm$^{-3}$, n(e$^{-}$)=n(H$^{+}$)=0.1 cm$^{-3}$, T=6,900 K and a relative Sun-cloud velocity of 25.8$\\pm$0.8 km s$^{-1}$. However, radiative transfer considerations in the LIC suggest that the quoted neutral density is a lower limit. Ulysses and EUVE observations of He$^{\\circ}$ indicate a cloud temperature of T=6,400 K. The magnetic field strength is weakly constrained to be in the range of 2--3 $\\mu$G. Models of the Ly$\\alpha$ absorption line towards $\\alpha$ Cen are consistent with an Alfven velocity of 20.9 km s$^{-1}$, which in turn is consistent with an interstellar magnetic field of 3 $\\mu$G in the absence of additional unknown contributions to the interstellar pressure. Ulysses and EUVE observations of interstellar He$^{\\circ}$ within the solar system give an upwind direction for the ``wind'' of interstellar gas through the solar system, in the rest frame of the Sun, of V=--25.9$\\pm$0.6 km s$^{-1}$ arriving from the galactic direction l=4.0$^{\\circ}$$\\pm$0.2$^{\\circ}$, b=15.4$^{\\circ}$$\\pm$0.6$^{\\circ}$. Removing solar motion from this vector gives an upwind direction for the LIC cloud in the LSR of V=--18.7$\\pm0.6$ km s$^{-1}$ arriving from the direction l=327.3$^{\\circ}$$\\pm$1.4$^{\\circ}$, b=0.3$^{\\circ}$$\\pm$1.0$^{\\circ}$. Through a combination of observations and theory, uncertainties in the LIC electron density are narrowing. Radiative transfer in the sightlines towards nearby stars require that cloud models must be combined with data in order to deduce properties at the cloud location. Radiative transfer models of ionization in the LISM show interesting results, but additional understanding of the input radiation fields is needed. The Local Fluff complex is structured and inhomogeneous. Striking progress would be made in understanding this structure if interstellar absorption lines could be observed at resolutions of $\\sim$1 km s$^{-1}$ in the ultraviolet. The most glaring uncertainty is the absence of detailed knowledge about the interstellar magnetic field. Many of the most abundant elements in the LIC are ionized, and densities of neutral atoms with FIPs less than 13.6 eV are typically down by 1--3 orders of magnitude from the dominant ions. The current approach of trying to understand the interaction of the ISM with the heliopause, from both the outside in and the inside out, is finally bearing fruit." }, "9804/astro-ph9804192_arXiv.txt": { "abstract": "The nuclei of a wide class of active galaxies emit broad emission lines with widths at half maximum (FWHM) in the range $10^{3}-10^{4}$ km s$^{-1}$. This spread of widths is not solely a consequence of the range of the luminosities of these sources since a plot of width versus luminosity shows a large scatter. We propose that the broad line emission region (BLR) is axially symmetric and that this scatter in line width arises from an additional dependence on the angle of the line of sight to the axis of the emission region. Such a relation is natural in unified models of active nuclei which link a variety of observed properties to viewing angle. Adopting a simple form for the line width as a function of luminosity and angle, and convolving this with the observed luminosity function, allows us to predict a line width distribution consistent with the available data. Furthermore, we use the relation between the equivalent width of a line and the luminosity in the continuum (the `Baldwin Effect') to predict an observed correlation between line width and equivalent width. The scatter on this correlation is again provided by angular dependence. The results have applications as diagnostics of models of the broad line emission region and in cosmology. ", "introduction": "In unified models of active galactic nuclei with spherically symmetric BLR the width distribution of the broad emission lines cannot be accounted for by luminosity dependence alone. Plots of line width versus continuum luminosity have a large scatter and show no significant correlation \\cite{W93,P97}. There is, however, growing evidence that the broad line region (BLR) is not spherical, but axisymmetric. \\begin{enumerate} \\item Observed samples of AGN \\cite{Wills86,W93,Brotherton94} suggest relations between line widths and R, the ratio of core to lobe dominance. Other samples \\cite{P97} find relations between line widths and $\\alpha_{\\rm ox}$, the continuum slope parameter from the optical to X--ray bands. Both of these parameters have some viewing angle dependence. \\item The continuum and line light curves of some active nuclei, eg 3C390.3 \\cite{Wamsteker97}, are most naturally interpreted in terms of a disc-like line emission region. It has also been suggested that some double peaked line profiles arise from discs (eg Arp 102b) although the interpretation in these cases is not so clear when time variability is taken into account. \\item Several axisymmetric disc-wind models, such as those of Cassidy \\& Raine \\shortcite{Cassidy96}, Chaing \\& Murray \\shortcite{Chaing96} and Emmering, Blandford \\& Shlosman \\shortcite{Emmering92} have been proposed and models of this type are gaining support from evidence for winds \\cite{Pasadena}. These will naturally predict some viewing angle dependence of line width. \\end{enumerate} It should be noted that Osterbrock \\shortcite{Osterbrock77} showed that a deficit of systems with narrower lines ruled out pure disc models, but such objections do not necessarily apply to axisymmetric models in general. In this paper we shall adopt a simple dependence of line width on both viewing angle and luminosity. Then: \\begin{enumerate} \\item We obtain a reasonable fit to the distribution of line widths. \\item Given the Balwin relation between line and continuum luminosity we predict a relation between line width and equivalent width compatible with the observed trend. The scatter on this relation is attributed to angular dependence. \\item We discuss how the width distribution can be used to test models of the BLR. \\item If the BLR is indeed axisymmetric, we show how the line width distribution can be used, in principle, to determine cosmological parameters. \\end{enumerate} ", "conclusions": "We conclude that the simple picture we have presented here accounts for the scatter in FWHM versus luminosity, accounts for the distribution of FWHM, and relates the trend in the EW with FWHM to the Baldwin relation. This may be useful as a diagnostic tool in discriminating between disc-wind models. The analysis has applications as a cosmological tool particularly as measurement of line widths is independent of any cosmological model." }, "9804/astro-ph9804317_arXiv.txt": { "abstract": "Collimated outflows from Young Stellar Objects (YSOs) can be seen as tracers of the accretion powered systems which drive them. In this paper I review some theoretical and observational aspects of YSO outflows through the prism of questions relating to the protostellar source. The issue I address is: can collimated outflows be used as ``fossils'' allowing the history of protostellar evolution to be recovered? Answering this question relies on accurately identifying where theoretical tools and observational diagnostics converge to provide unique solutions of the protostellar physics. I discuss potential links between outflow and source including the time and direction variability of jets, the jet/molecular outflow connection, and the the effect of magnetic fields. I also discuss models of the jet/outflow collimation mechanism. ", "introduction": " ", "conclusions": "The issues cited in this paper are associated with outflows. How do these issues specifically relate to questions inherent to the physics of accretion? The time variability of jets relates to the time-dependence of accreation, the FU Ori outbursts being a notable example. The direction variability of jets relates to the global dynamics and stability of accretion disks. Livio \\& Pringle 1997, for example, have shown that radiation induced warping of disks may lead to precession in magneto-centrifugal jets. The presence and structure of magnetic forces in jets relates to the existence and form of large scale fields in the disks. If nose-cones do not occur in real YSO jets then perhaps mechanisms which rely on strong toroidal fields are excluded. Thus YSO jets and outflows offer a unique opportunity for the study of accretion powered systems. Protostellar outflows can be observed with exquiste detail in a variety of wavelengths including diagnostic spectral lines. The quality of the data combined with the long lookback time inherent to the outflows offers the possibilty that a large fraction of individual protostar's history might be recovered if we learn were and how to look. We are a long way from this now but the prospect of having such capabilities is very exciting." }, "9804/astro-ph9804121_arXiv.txt": { "abstract": " ", "introduction": "A long way has been run from the first views developped to explain the formation of galaxies. In 1962, Eggen, Lynden-Bell \\& Sandage designed the collapse scenario, where all galaxies are created with their morphological type, according to their angular momentum. Their potentials remained axisymmetric, so that no angular momentum could be redistributed through gravity torques; the total mass and gas content was already there at first collapse. For elliptical galaxies, the violent/single collapse picture still remains in some modified form, although the most developped and adopted scenario is through agglomeration of a large number of clumps (e.g. van Albada 1982, Aguilar \\& Merritt 1990), that produces de Vaucouleurs profiles in $r^{1/4}$. The merger picture (Toomre 1977, Schweizer 1990), where ellipticals are formed by progressive interaction and coalescence of many parent galaxies, is favored in hierarchical cosmogonies. For spiral galaxies, the scenario involves now much more internal dynamical evolution. Due to gas dissipation and cooling, gravitational instabilities are continuously maintained in spiral disks, and they drive evolution in much less than a Hubble time. Spiral galaxies are open systems, that accrete mass regularly, and their morphological type evolves along the Hubble sequence. Non-axisymmetric perturbations, such as bars or spirals, produce gravity torques that drive efficient radial mass flows; vertical resonances thicken disks and form bulges, and the mass central concentration can destroy bars. Accretion of small companions can also disperse bars and enlarge the bulge. A major merger can destroy disks entirely and form an elliptical. The first role of galaxy interactions is to trigger internal evolution, that we will consider now, in the next section. Specific aspects of galaxy interactions and mergers will then be detailed in section \\ref{envir}. ", "conclusions": "" }, "9804/astro-ph9804298_arXiv.txt": { "abstract": "We have produced radio maps, using the ATCA, of the central regions of six southern Seyfert 2 galaxies (NGC 1365, 4945, 6221, 6810, 7582, and Circinus) with circumnuclear star formation, to estimate the relative contribution of star formation activity compared to activity from the active galactic nucleus (AGN). The radio morphologies range from extended diffuse structures to compact nuclear emission, with no evidence, even in the relatively compact sources, for synchrotron self--absorption. In each case the radio to far--infrared (FIR) ratio has a value consistent with star formation, and in all but one case the radio to [FeII] ratio is also consistent with star formation. We derive supernova rates and conclude that, despite the presence of a Seyfert nucleus in these galaxies, the radio, FIR, and [FeII] line emission are dominated by processes associated with the circumnuclear star formation (i.e. supernova remnants and H~II regions) rather than with the AGN. ", "introduction": "Circumnuclear star formation is common in Seyfert galaxies, but the relationship between the Seyfert nucleus and the surrounding star formation is not well understood (see Filippenko 1992), and both evolutionary and causal relationships have been suggested. For example, a nuclear starburst may develop into a massive star cluster or black hole (Norman \\& Scoville 1988), or conversely the energy output from an active galactic nucleus (AGN) may trigger circumnuclear star formation (Sanders \\& Bania 1976). More recently, studies have indicated that star formation is occurring in and around the torus (e.g. Cid Fernandes \\& Terlevich 1992; Davies \\etal 1997) that is thought to dictate the type of Seyfert nucleus we observe. Dust is a common feature of the circumnuclear environment of active galaxies. At optical wavelengths, it can obscure our view of the nucleus and hide any evidence of an AGN. However, radio observations are not only unaffected by dust but also have the advantage of high spatial resolution, and can be important in understanding the role of the various processes in active galaxies. For example, starburst galaxies generally have diffuse radio morphologies dominated by synchrotron emission from cosmic rays accelerated by supernovae, while Seyfert galaxies sometimes exhibit well--collimated radio jets and a compact nuclear radio source (see review by Condon 1992). High-resolution radio synthesis images, with sub-arcsec resolution, have found compact sources at the nucleus, and extended emission consisting of radio jets, ouflows, diffuse emission, and discrete sources such as HII regions and SNRs (e.g. Ulvestad \\& Wilson 1984; Kronberg \\etal 1985; Antonucci \\& Ulvestad 1988; Carral, Turner \\& Ho 1990; Condon \\etal 1991; Forbes \\etal 1994; Collison \\etal 1994; Sandqvist \\etal 1995). High spatial resolution, multi--frequency radio observations can also be used to test the idea of advection-dominated accretion flows around black holes. For such flows the radio emission depends strongly on the mass of the black hole and is characterised by an inverted radio spectrum (Yi \\& Boughn 1997). Here we present 3 and 6cm radio continuum images from the Australia Telescope Compact Array (ATCA) of 6 such galaxies which show evidence for narrow high excitation optical forbidden lines, and are classified as Seyfert 2s. In the case of NGC 1365, broad lines have also been detected from the nuclear region. All six galaxies show evidence for circumnuclear star formation, and several are barred. We discuss the radio morphologies of these galaxies and possible emission mechanisms, and we compare the radio data with that from other wavelengths to assess the relative contributions to the radio flux from star formation and the AGN. ", "conclusions": "In Figures 1 to 12 we show the 3 and 6cm radio images for the six galaxies in our sample. These galaxies do not represent a complete sample in any sense, but rather were chosen as relatively well--known southern Seyfert 2 galaxies that generally lacked high resolution radio maps. Our radio data for the Circinus galaxy have been presented elsewhere along with near--infrared line imaging (Davies \\etal 1997), although we include it here for comparison purposes with the other galaxies. The images show a variety of radio morphologies which range from Circinus, with a strong, compact nucleus, to NGC 1365, with a extended region of discrete sources or hot-spots. The beam shape is shown in the lower left of each image. Care is needed when interpreting these images, as (a) our observations are optimised for studying the nuclear region, and so our images will not correctly represent the extended diffuse emission from the disk of the galaxy, and (b) most of the galaxies have high inclinations so that confusion effects may play some role in the observed radio morphology. Flux measurements in 2 arcsec and 6 arcsec diameter apertures for each galaxy are given in Table 2. We also list the 6cm to 3cm spectral index after smoothing the 3cm image to match the 6cm resolution. The spectral indices (F $\\propto \\nu ^{\\alpha}$) range from flat ($\\alpha$ $\\sim$ 0) to steep ($\\alpha$ $\\sim$ --1). The total 6cm flux in Table 2 is given both for our images and for the single--dish observations by Wright \\etal (1994, 1996). The difference between these indicates the amount of diffuse emission which is missing from our images. We also calculated the maximum brightness temperatures for each image (both 3 and 6 cm) and the maximum value is given for each source. In Table 3 we list various derived quantities for the sample including Hubble type, distance, inclination, 6cm and [FeII] line luminosities and the SN rate. The [FeII] measurements are from Moorwood \\& Oliva (1988) in a 6 arcsec diameter aperture, except for the Circinus galaxy in which we use a nuclear [FeII] flux from Davies \\etal (1997). None of the [FeII] fluxes have been corrected for extinction. The SN rate is calculated separately from both the 6cm and [FeII] line flux in the 6 arcsec aperture (corresponding to $\\sim$ 500 pc at a typical distance of 20 Mpc). The 6cm SN rate is calculated assuming that all of the 6cm flux is non--thermal emission from cosmic rays accelerated by SNRs (e.g. Condon \\& Yin 1990). This may give an overestimate of the rate because it ignores any nuclear flux (which may be significant) and the contribution from thermal emission (which is unlikely to be significant). For the [FeII] SN rate we simply assume an average luminosity of 2 $\\times$ 10$^{36}$ erg s$^{-1}$ over an adiabatic lifetime of 2 $\\times$ 10$^{4}$ yrs (e.g. Norris \\& Forbes 1995). For most galaxies the two methods give rates within a factor of two, the notable exception being NGC 4945 (which is discussed further below). \\subsection{Radio Spectral Indices} The radio spectral index of Seyfert and starburst galaxies is determined by four mechanisms. 1. Cosmic rays within the galaxy are generated and re--accelerated by supernovae and supernova remnants, and then interact with the interstellar magnetic field to emit synchrotron emission. This synchrotron emission, with a typical spectral index of $\\alpha$ $\\sim$ --0.7, is expected to dominate the radio power of starburst galaxies, and should appear as a diffuse component in radio images of these sources. 2. Relativistic particles ejected from the massive black hole at the nucleus of a galaxy may generate intense synchrotron emission, similar to that seen in radio--loud galaxies and quasars. The cores in these radio-loud objects typically have a flat--spectrum core, indicating synchrotron self--absorption, and steep--spectrum extended radio--lobes, suggesting cooling of high--energy electrons. However, synchrotron self--absorption is important only for brightness temperatures greater than 10$^{10}$ K (Condon 1992). Most Seyfert galaxies, on the other hand, are observed to have much lower brightness temperatures than this in the core, so that synchrotron self-absorption is not significant in these sources. This is confirmed by the observed core spectral indices, which are frequently in the region of --0.7. 3. When the radiative efficiency in the accretion disk is low, an advection-dominated accretion flow (ADAF) may operate. The radio emission in these sources is dominated by synchrotron emission from a hot plasma, and the emission from such flows is predicted to give rise to inverted spectra with typical indices of +0.4 (Yi \\& Boughn 1997). The ADAF radio emission mechanism has only recently been proposed and should be regarded as untested at this stage. The data here are unlikely to provide a definitive test because of insufficient resolution, and all radio spectral indices measured here are negative. We will therefore not consider this mechanism further, except to note that this mechanism would be indicated by inverted-spectrum emission from a low-brightness-temperature core. No source here falls into this category. 4. H II regions in our galaxy generate free--free emission from hot electrons. Most are optically thin, giving a flat spectrum, although some compact H II regions become optically thick at centimetre wavelengths, giving a spectral index $\\sim$+2. However, the integrated flux of such regions is generally insignificant compared to the synchrotron emission of the host galaxy. 5. The radio emission from ultra--luminous infrared galaxies is optically thick to free--free absorption, so that the typical synchrotron spectrum of these galaxies is flattened at low frequencies (Condon et al. 1991). The combined result of these effects in Seyfert and starburst galaxies is to produce a typical radio spectral index of --0.7 (from the extended synchrotron emission) with a flattening at low frequencies in some starburst sources because of free--free absorption. Table 2 shows that the nuclear spectral indices of three of the galaxies (NGC 1365, NGC 6221, NGC 7582) is --0.5 or steeper on both the 2 arcsec and the 6 arcsec scale, showing evidence for neither free--free nor synchrotron absorption. In the other three galaxies, the cores have flatter spectra, but the brightness temperatures ($\\le$ 8300 K) are too low for synchrotron self--absorption, indicating that free--free absorption is responsible for the flattening. Of course, we cannot rule out the presence of a weak synchrotron self--absorbed core in the nucleus of any of these galaxies. However, comparison of the radio fluxes in a 2--arcsec aperture with the flux in a 6--arcsec aperture in Table 2 shows that the luminosity of any such core is small compared to the surrounding emission. Therefore any such core does not contribute significantly to the overall energy budget of the nuclear region of the galaxy, and is not responsible for the overall flat spectrum of the nuclear region.. This degree of free-free absorption flattening indicates either a high star formation rate (Condon \\etal 1991) or that we are viewing the AGN through optically--thick narrow-line-region clouds (Roy \\etal 1994). \\subsection{The Radio - [FeII] Correlation} Forbes \\& Ward (1993) discovered that the 6cm radio emission in the central regions of active galaxies is strongly correlated with the near--infrared [FeII] 1.64$\\mu$m line emission. This relation exists over several orders of magnitude. With a larger sample, Simpson \\etal (1996) were able to show that Seyfert and starburst galaxies follow slightly different radio--[FeII] relations. For starburst galaxies the relation, with slope $\\sim$ 1, can be reasonably explained by SNRs which are responsible for both the non--thermal radio emission and the fast shocks that provide the [FeII] excitation. However, the situation for Seyfert galaxies (which reveal a correlation slope of $\\sim$ 0.7) is less clear. Simpson \\etal argued that photo-ionisation from the Seyfert nucleus can cause this relationship, with a contribution from radio--jet induced shocks in some cases. In Fig. 13 we show the [FeII]/6cm ratio for our sample galaxies, compared with the Seyfert and starburst relations of Simpson \\etal (1996). The 1$\\sigma$ dispersion of galaxies about the relations is $\\sim$ $10^{0.5}$. For the Seyferts studied here, we find a large degree of star--formation activity compared to photo--ionisation from an AGN, and so we might expect them to lie closer to the starburst relation than the Seyfert one. This indeed appears to be the case for four galaxies, although one (NGC 7582) is closer to the Seyfert relation and NGC 4945 falls well away from either relation. The [FeII]/6cm ratio of NGC 4945 is about a factor of 100 lower than typical active galaxies, and we discuss this further in Section 3.4 below. We note however that given the dispersion in the relations, and the low luminosities ($\\le$ 10$^{40}$ erg s$^{-1}$) of the galaxies studied here, this is not a sensitive test of the excitation mechanism. \\subsection {The Radio -- FIR Correlation} Normal spiral and starburst galaxies show a tight correlation between their radio and FIR luminosity (e.g. Wunderlich et al. 1987). This correlation, which extends over five orders of magnitude, is true for both flux density and luminosity, and cannot be accounted for by selection effects, or by a simple ``richness effect''. While a detailed mechanism to explain this correlation has yet to be established, it is almost certainly the result of star formation, which generates both the synchrotron radio emission and the thermal FIR emission. This is supported by the fact that all objects that are dominated by star formation (HII galaxies, normal spirals, starburst galaxies) do follow the correlation. On the other hand, Sopp \\& Alexander (1991) showed that radio--loud quasars and radio galaxies clearly do not follow the radio--FIR correlation. Thus whether or not a galaxy follows this correlation may be used as an indicator of the dominant radio luminosity source of the galaxy. Norris \\etal (1988) and Roy \\etal (1997) showed that Seyfert galaxies, unlike radio--loud quasars, do roughly follow the radio--FIR correlation, but with a looser fit than normal spirals and starbursts. This suggests that the bolometric luminosity of Seyfert galaxies may be dominated by star formation. This is supported by off--nuclear optical and infrared observations of Seyferts, which show the same line ratios and luminosities as starburst galaxies (Bransford et al. 1997). Thus, although the nuclear optical spectra of Seyfert galaxies are clearly dominated by an AGN, in many cases the integrated radio emission and the FIR emission are dominated not by the AGN but by circumnuclear star formation. The degree to which an individual galaxy follows this correlation is most conveniently expressed by the parameter q -- the logarithm of the FIR to radio ratio. The conventional definition of q follows that of Helou \\etal (1985), who define q in terms of the 1.49 GHz radio flux. For our purposes, we adapt Helou's definition to our observing frequency of 4.8 GHz by assuming a spectral index of --0.7, and therefore define it as \\hspace{.5in}q $\\equiv$ log[(FIR/3.75 x 10$^{12}$ Hz)/(2.26 x S$_{\\rm 4.8 GHz}$]\\hspace{1.5in}(1) \\hspace{.5in}where FIR $\\equiv$ 1.26 x 10$^{-14}$(2.58S$_{\\rm 60\\mu}$+S$_{\\rm 100\\mu}$)\\hspace{2in}(2) Typical values of q from the IRAS Bright Galaxy Sample are 2.34 for normal spirals, 2.21 for starburst galaxies, and less than 2 for radio--loud AGNs (Condon \\etal 1991). All the galaxies studied here except NGC 4945 have q in the range 2.2 to 2.3, which places them firmly in the middle of the radio--FIR correlation, and suggests that most of their radio and FIR luminosity is produced by star formation. We discuss the case of NGC 4945 (q = 1.88) below. \\subsection{Individual Galaxies} Here we discuss each galaxy in turn, starting with an extended discussion of NGC 4945. To avoid repetition, we note that in every case other than NGC 4945, the spectral index, radio--FIR ratio, and [FeII]--radio ratio are all consistent with star formation, rather than an AGN, being the dominant source of radio emission. \\noindent {\\bf NGC 4945} This infrared luminous galaxy is nearly edge--on and is located in a nearby loose group. Although we list it as a barred galaxy in Table 3, there is a continuing debate about the reality of the bar (e.g. Harnett \\etal 1989). Koornneef (1993) described NGC 4945 as a post--starburst galaxy with no evidence for an AGN. However Moorwood \\& Oliva (1994) have argued that the central regions do show signs of ongoing young star formation. Evidence for a heavily obscured AGN now come from the variable hard X--rays (Iwasawa \\etal 1993), and the presence of a compact radio core in VLBI observations (Sadler \\etal 1995). The galaxy contains a thick torus or ring with a radius of $\\sim$ 150 pc (Koornneef 1993; Moorwood \\etal 1996). Harnett \\etal (1989) found that the radio emission has a strong central contribution with emission extended 10 arcmin perpendicular to the major axis. Multi--frequency observations have been carried out by Elmouttie \\etal (1997). They focused on the large scale structure using a beam size of $\\sim$ 25$^{''}$, and found that the spectral index steepens from the central region to the main disk of the galaxy. Furthermore, NGC 4945 is notable for the fact that it is one of the few galaxies (along with Circinus) known to contain water megamasers. Such megamasers have been cited in NGC 4258 (Miyoshi \\etal 1995) as the strongest evidence known for a massive black hole in an AGN. Preliminary VLBI imaging (Greenhill \\etal 1997) of the megamasers supports the model that they are in a Keplerian disk surrounding the black hole. Our radio image is dominated by strong nuclear emission, and emission extended along the disk of the galaxy. However there is also evidence of some filamentary structure perpendicular to the major axis. Such extended emission may be associated with the outflowing superwind in this galaxy (Nakai \\etal 1989; Lipari, Tsvetanov \\& Macchetto 1997). The extended emission has a steeper spectral index ($\\alpha \\sim -0.8$) than the radio nucleus, however we note that our data are less sensitive to extended emission (particularly at 3cm) which makes the spectral index somewhat less certain. The nucleus has a relatively flat spectral index of $\\alpha$ = --0.3. The brightness temperature in the central 2 arcsec of this source (i.e. 7000 K) is still far too low for synchrotron self--absorption, indicating that the star formation activity is particularly intense, to provide the necessary free--free absorption. The obscuration inferred from the X--ray data suggest that the extinction towards the AGN could be as high as A$_V$ $\\sim$ 2500. We noted above (in Section 3.3) that the radio--FIR ratio for this galaxy is unusually low (i.e. q = 1.88), which at first sight appears to suggest that an AGN is responsible for much of the radio emission. However, this galaxy is so near that not all the FIR flux was in the single IRAS aperture, and so the IRAS flux listed in the IRAS Point Source Catalog may be an underestimate. Rice \\etal (1988) have estimated the total FIR flux by co--adding IRAS images and obtain a higher value for the FIR fluxes, which raises the value of q to 2.1, suggesting that the radio emission in this galaxy is again dominated by star formation rather than by an AGN. An interesting property of NGC 4945 is that the ratio of the [FeII] line luminosity to 6cm radio emission is only 0.63, which is almost a factor of 100 less than is typical for active galaxies (see Fig. 13). We now consider a number of possible reasons for this. \\begin{itemize} \\item The reduced ratio could be due to extinction (by dust) of the [FeII]. However, Moorwood \\& Oliva (1994) estimate that the extinction in the [FeII] line emitting zone is 1.8 mag or a factor of five, which is insufficient to produce the observed effect. \\item It could be because of nuclear radio emission from an AGN which is not accompanied by [FeII] line emission. We have shown above that the radio/FIR ratio for the galaxy as a whole is consistent with star formation activity. However, the central 6 arcsec (over which we measure the [FeII]/6cm ratio) contributes only 9\\% of the total radio flux, and we have no information on the radio/FIR ratio in the nucleus, so the radio flux from the AGN could be abnormally large. In this case, we would expect the [FeII]/6cm ratio to approach the usual value as we increase the area over which we integrate the flux (for both 6 cm and [FeII]). However, Moorwood \\& Oliva (1994) quote a total [FeII] flux over an emitting region of 18 $\\times$ 21 arcsec to be 12 $\\times$ 10$^{-14}$ erg s$^{-1}$ cm$^{-2}$, or an observed log luminosity of 38.81 erg s$^{-1}$. The 6cm radio luminosity over a similar area is 40.21 erg s$^{-1}$, giving a [FeII]/6cm ratio of 0.03, which is even lower than the value in the nucleus, indicating that the [FeII]/6cm ratio falls off with distance from the galaxy centre, and that the low value is not a consequence of nuclear radio emission. \\end {itemize} Thus the abnormally low [FeII]/6cm ratio in NGC 4945 of 0.63 is produced in the region surrounding the nucleus, where the radio emission (with a spectral index of $\\sim$ --0.7) is due to SNRs in the galaxy disk, the outflowing starburst superwind discussed above, or perhaps a radio jet. Pure SNRs produce ratios of about 500 i.e. well in excess of typical galaxy values, so this would tend to give an enhanced ratio. In a $6^{''} \\times 6^{''}$ aperture, the superwind galaxies M82 and NGC 253 have [FeII]/6cm ratios of 75 and 52 respectively, although the superwind itself in NGC 253 does not seem to produce significant [FeII] line emission (Forbes \\etal 1993). Again such ratios are significantly higher than seen in NGC 4945. The data for Seyfert galaxies with clear radio jets are limited. For NGC 4151 and NGC 1068 the measured ratios are 28 and 9. This is closer to the NGC 4945 value but still a factor of at least 10 too high. We conclude that the abnormal [FeII]/6cm ratio in NGC 4945 is due to either (a) a starburst superwind, which produces substantial radio emission but very little [FeII] line flux (due perhaps to an unknown excitation effect or low density in the wind), or (b) a radio jet which dominates the extended radio emission on the few--arcsec scale but which does not produce significant [FeII] emission. \\noindent {\\bf NGC 1365} This is a well--studied barred galaxy. The central region reveals broad and narrow emission lines (Veron \\etal 1980) surrounded by a circumnuclear ring of star formation (Edmunds \\& Pagel 1982; Saikia \\etal 1994). The star formation, combined with the obscuring effects of dust, give the appearance of a hot-spot nucleus (Sersic \\& Pastoriza 1965). A high excitation outflow from the nucleus has been seen (e.g. Hjelm \\& Lindblad 1996). High resolution radio continuum observations have been reported by several workers (e.g. Sandqvist, Jorsater \\& Lindblad 1982, 1995). In particular, Sandqvist, Jorsater \\& Lindblad (1995) observed it with the VLA at 20, 6, and 2 cm. Their radio images revealed a weak nucleus surrounded by a elongated $\\sim$ 8 $\\times$ 20 arcsec (a/b = 0.4) ring of hot-spots, or components. They labelled a number of components A--H, of which B is blended with A and C is blended with D at $\\ge$ 1 arcsec resolutions. Our radio image, shown in Fig. 1, is consistent with theirs, except that we identify one additional component to the SW, which we call `J'. The hot-spots generally have steep spectra with 6cm luminosities of $\\sim$ 10$^{36}$ erg s$^{-1}$ which suggests that the radio emission from each component is made up of several SNRs. A combined radio and X--ray study of the nucleus and surrounding regions has been carried out by Stevens, Forbes \\& Norris (1998). The radio nucleus does not appear to have an X--ray counterpart. Furthermore the X--ray emission is consistent with star formation processes. Stevens \\etal conclude that if NGC 1365 harbours a black hole it is largely inactive. \\noindent {\\bf NGC 6221} Located in a small group, NGC 6221, is a barred galaxy with a weak Seyfert nucleus. The galaxy may be interacting with NGC 6215 and has a nuclear bar (Koribalski 1996). The radio emission from NGC 6221 is extended in an symmetric bar--like structure. The spectral index of the nucleus and bar are non--thermal with $\\alpha$ $\\sim$ --0.6, indicative of SNRs. The radio morphology and other properties are all consistent with star formation being the dominant source of radio emission. \\noindent {\\bf NGC 6810} This early type spiral is the most distant in our sample and has not been well studied to date. It may contain a bar and ring structure (Buta 1995), and does not appear to have been imaged before at radio wavelengths. Our radio image reveals a dominant nucleus surrounded by diffuse extended radio emission. Both the nucleus and surrounding region have flat spectral indices, but the brightness temperature is too low for this to be attributable to synchrotron self--absorption, which suggests that the radio spectrum is flattened by free--free absorption from young star formation. The radio morphology and other radio properties are all consistent with star formation being the dominant source of radio emission. Interestingly, recent high resolution optical spectra do not confirm the status of NGC 6810 as a Seyfert galaxy (Heisler 1998), thus it appears to have been misclassified. \\noindent {\\bf NGC 7582} This narrow line X--ray galaxy is located in the Grus loose group along with NGC 7590 (a Seyfert 2; Ward \\etal 1980), NGC 7552 (a starburst; Forbes \\etal 1994) and NGC 7599. Several HI bridges connect group members (Koribalski 1996). Morris \\etal (1985) provide evidence for both a rapidly--rotating $\\sim$ 1 kpc ring of circumnuclear star formation and high excitation gas moving outwards from the nucleus. Ulvestad \\& Wilson (1984) imaged the galaxy at 6 cm using the VLA with a beam size of $\\sim$ 1.5 arcsec. They measured a total 6cm flux of 69 mJy. We find a linear, double--peaked morphology to the radio emission. The southern peak appears to lie at the centre of the outer radio isophotes and is presumably the true nucleus. The nucleus has a steep spectral index ($\\alpha$ = --0.7) indicating non--thermal emission. To the NW by $\\sim$ 3 arcsec lies a second peak, which could be a second nucleus. However, it lies roughly along the bar/major axis position angle (P.A. $\\sim$ 150$^{\\circ}$). It has a 2 arcsec diameter 6cm flux of 10 mJy and a spectral index of --0.9. This second peak could therefore be a radio jet or simply a discrete star formation region occurring along the galaxy bar. The inferred SN rate in the central 6 arcsec (870 pc) is the highest in our sample (except possibly for NGC 4945) at about 1 SN every 8 years. Although the radio morphology suggests a linear Seyfert jet, the spectral index, radio--FIR ratio, and [FeII]--radio ratio are all consistent with star formation being the dominant source of radio emission. \\noindent {\\bf Circinus} The Circinus galaxy is perhaps the closest Seyfert galaxy known but is difficult to observe due to its proximity to the Galactic plane and large internal obscuration. Confirmation of an AGN comes from the the presence of high excitation coronal lines (Oliva \\etal 1994), X--ray emission (Matt \\etal 1996), and a compact radio core (Heisler \\etal 1998). Like NGC 4945, Circinus is one of the few water megamaser galaxies. The megamasers in Circinus are stronger but less extreme than those in NGC 4258, and have the curious property of fluctuating on a time scale of minutes (Greenhill \\etal 1997), indicating a particularly compact source. Preliminary VLBI imaging of the megamasers (Ellingsen \\etal 1998) indicates that the maser region is extended with a velocity gradient aligned with that of the parent galaxy, and perpendicular to the jet. We regard this as strong evidence for a massive black hole in this galaxy. Marconi \\etal (1994) found both a circumnuclear starburst ring and an ionisation cone. The [OIII] ionisation cone is asymmetric extending only to the NW, with some high excitation lines also seen in the cone region. They estimated the extinction to the nucleus to be A$_V$ $\\sim$ 20. The HI gas distribution shows spiral arms, a bar and a central `hole' (Koribalski 1996, Jones \\etal 1998). High resolution observations of the central region indicate a rapidly--rotating gas ring with a diameter of $\\sim$ 400 pc (Koribalski 1996). Observations with the ATCA have been carried out at 13 and 20 cm by Elmouttie \\etal (1995). They found extended radio lobes perpendicular to the galaxy major axis (position angle = 30$^{\\circ}$) with a spectral index of $\\alpha \\sim$ --0.7. Our 3 and 6cm radio images, observed as part of this project, have been published, along with near--infrared line images, by Davies \\etal (1997). The radio data indicate that the nucleus is marginally resolved with a flat spectral index. The low brightness temperature indicates that this is due to free--free absorption rather than synchrotron self--absorption in a compact AGN source. There are also faint hints of extended emission which may be associated with outflowing material. Despite the clear indication of a compact AGN in the radio images, the other radio indicators are consistent with the more extended radio emission being dominated by star formation activity.\\\\" }, "9804/astro-ph9804251_arXiv.txt": { "abstract": "The innermost regions of quasars can be resolved by a gravitational-lens {\\lq}telescope{\\rq} on scales down to a few AU. For the purpose, X-ray observations are most preferable, because X-rays originating from the innermost regions, can be selectively amplified by microlensing due to the so-called `caustic crossing'. If detected, X-ray variations will constrain the size of the X-ray emitting region down to a few AU. The maximum attainable resolution depends mainly on the monitoring intervals of lens events, which should be much shorter than the crossing time. On the basis of this idea, we performe numerical simulations of microlensing of an optically-thick, standard-type disk as well as an optically-thin, advection-dominated accretion flow (ADAF). Calculated spectral variations and light curves show distinct behaviors, depending on the photon energy. X-ray radiation which is produced in optically thin region, exhibits intensity variation over a few tens of days. In contrast, optical-UV fluxes, which are likely to come from optically thick region, exhibit more gradual light changes, which is consistent with the microlensing events so far observed in Q2237+0305. Currently, Q2237+0305 is being monitored in the optical range at Apache Point Observatory. Simultaneous multi-wavelength observations by X-ray sattelites (e.g., ASCA, AXAF, XMM) as well as HST at the moment of a microlens event enable us to reveal an AU scale structure of the central accretion disk around black hole. ", "introduction": "The high power output from quasars is usually attributed to the combination of a supermassive black hole with a surrounding accretion disk. This belief is supported by a number of observations that indicate the presence of a deep gravitational potential well or a hot gas disk at the center of quasars or other active galactic nuclei; e.g., measurements of stellar velocity dispersion clearly showed a peculiar increase toward the center (Young et al. 1978; Sargent et al. 1978; see also Ford et al. 1994; Harms et al. 1994). Malkan (1983) found that the optical to UV spectra are well fitted by the standard-type accretion disk model (Shakura \\& Sunyaev 1973). Recently, by far the best evidence of a supermassive black hole has been found by radio observations of nuclear H$_2$O maser sources in NGC4258 (Miyoshi et al. 1995). Alternatively, we can infer the presence of a relativistic object from the asymmetric Fe line profile (Tanaka et al. 1995). These observational results are all attractive, but still the real vicinity of a putative black hole has not been resolved. Q2237+0305 (e.g., Huchra et al. 1985) is the first object, in which the quasar microlensing events were detected (Corrigan et al. 1987; Houde \\& Racine 1994; see also Ostensen et al. 1996). These observations suggest that microlensing events take place roughly once per year. This rather high frequency is consistent with the microlens optical depth of $\\tau \\sim 0.8$ obtained by the realistic simulation of the lensing galaxy (i.e., Wambsganss \\& Paczy\\'nski 1994). We consider, here, specifically the microlensing events of this source caused by the so-called `caustic crossings' (see Yonehara et al. 1997 for single-lens calculations). Several authors have already analyzed this `caustic' case based on a simple model for quasar accretion disk (e.g., Wambsganss \\& Paczy\\'nski 1991; Jaroszy\\'nski, Wambsganss \\& Paczy\\'nski 1992). So far, however, only the standard-type disk, which is too cool to emit X-rays, has been considered, and thus the property of an X-ray microlensing of quasar, e.g., Q2237+0305, has not been predicted. We stress here the significance of X-ray observations to elucidate the physics of the innermost parts of the disk, since X-rays specifically originate from a deep potential well. The observations allow us to assess the extension of hot regions on several AU scales and resultantly to deduce the mass of a central massive black hole. In this $Letter$, we propose to investigate quasar central structure by using X-ray microlensing of Q2237+0305. In section 2, we describe the method for resolving X-ray emission properties of the inner disk structure on a scale down to a few AU. In section 3, we calculate the disk spectra and light curves during microlensing. We here use realistic disk models: the optically-thick, standard disk (Shakura \\& Sunyaev 1973) and the optically-thin, advection-dominated accretion flow (ADAF, Abramowicz et al. 1995; Narayan \\& Yi 1995; see also Ichimaru 1977). ", "conclusions": "" }, "9804/cond-mat9804137_arXiv.txt": { "abstract": "We study with Monte Carlo methods an ensemble of $c=-5$ gravity graphs, generated by coupling a conformal field theory with central charge $c=-5$ to two-dimensional quantum gravity. We measure the fractal properties of the ensemble, such as the string susceptibility exponent $\\gamma_s$ and the intrinsic fractal dimensions $d_H$. We find $\\gamma_s = -1.5(1)$ and $d_H = 3.36(4)$, in reasonable agreement with theoretical predictions. In addition, we study the critical behavior of an Ising model on a {\\it quenched} ensemble of the \\mbox{$c=-5$} graphs and show that it agrees, within numerical accuracy, with theoretical predictions for the critical behavior of an Ising model coupled {\\it dynamically} to two-dimensional quantum gravity, provided the total central charge of the matter sector is $c=-5$. From this we conjecture that the critical behavior of the Ising model is determined solely by the average fractal properties of the graphs, the coupling to the geometry not playing an important role. ", "introduction": "Randomness in statistical systems arises in a variety of situations and is a very rich and complex subject. Quenched randomness is frequently used in studying the role of impurities and inhomogeneities in real physical systems where the characteristic time-scale of the disorder is much longer than other dynamics of the system. Annealed randomness, on the other hand, arises naturally in studies of fluctuating geometries, such as two-dimensional quantum gravity or fluid membranes, where the disorder is dynamically modified by interaction between the geometry and matter fields living on the surfaces. For a statistical system coupled to random disorder, either in a quenched or annealed approach, the main question is to assess the effect randomness has on the critical behavior of the pure system. One prediction in this direction is the Harris conjecture \\cite{harris} which states that randomness changes the values of critical exponents only if the specific heat exponent $\\alpha$ of the pure system is positive. This conjecture has been studied in many models with quenched disorder, such as the $2d$ Ising model \\cite{dots} (where the Harris criterion is ambiguous as $\\alpha = 0$) and the Potts model \\cite{bondpott}. For both models a change in the critical behavior is observed. All the above mentioned studies deal with weak disorder. More recently the critical behavior of systems on lattices with fractal structure very different from a flat surface has been investigated. Such systems arise naturally when matter, in the form of conformal field theories, is coupled to two-dimensional quantum gravity. These models can be studied either in a continuum formulation, by Liouville field theory, or using discretized approaches like, for example, models of dynamical triangulations, formulated either as matrix models or studied with numerical simulations. For these systems the disorder is, however, different from the one discussed above in that it is annealed, i.e.\\ the models couple dynamically to fluctuations in the geometry. A remarkable degree of universality does emerge for models coupled to two-dimensional quantum gravity. Namely, the change in the critical behavior of the systems, and their back-reaction on the geometry, only depends on the total central charge of the matter sector. This manifests itself in the so-called KPZ scaling relation which describe how the scaling dimensions of conformal operators are changed by the interaction with gravity \\cite{kpz}. Moreover, this universality also extents to the fractal structure of the surfaces, from which we derive the string susceptibility exponent $\\gamma_s$ and the fractal dimension $d_H$. In view of this universality it is tempting to conjecture that the critical behavior of a particular system, when coupled to a fluctuating geometry, only depends on the (average) fractal structure of the surface. Details of the interaction between the system and the geometry, or the geometrical fluctuations, are not important as such --- they only serve the purpose of defining the average fractal geometry. If this conjecture is true it implies that how the average over disorder is performed, i.e.\\ that the disorder is annealed, is not essential. In particular, predictions of the KPZ scaling relation for the change in the critical behavior should just as well apply to models with quenched disorder, {\\it provided the quenched average is taken over the same ensemble of disorder as is generated in the annealed approach}. There are some recent simulations that have addressed the question of the critical behavior of spin models on a quenched ensemble of graphs generated by two-dimensional quantum gravity. Both the Ising model \\cite{bhj} and the 10-state Potts model \\cite{bjj} have been studied on an ensemble of pure gravity graphs ($c=0$). For the Ising model a critical behavior compatible with an Ising model coupled dynamically to gravity was found, although the accuracy of the results is not sufficient to rule out the conjecture discussed above. The goal of this paper is two-fold. First, we want to investigate the fractal geometry of two-dimensional quantum gravity coupled to a conformal field theory with central charge $c=-5$. More precisely, we want to determine the fractal dimension of the corresponding surfaces, using recently developed finite-size scaling methods \\cite{hausd,janhaus} and to compare it to the (contradictory) theoretical predictions that exist \\cite{anhaus1,anhaus2}. Second, we want to investigate the critical behavior of an Ising model on a quenched ensemble of $c=-5$ graphs and to compare it with predictions from Liouville theory, for the critical behavior of an Ising model coupled dynamically to two-dimensional quantum gravity, and to verify, or disprove, our conjecture about the effect of the disorder. Our motivation for choosing $c=-5$ is that both its predicted fractal structure and the critical behavior of the Ising model is substantially different from both a flat space and for a pure two-dimensional quantum gravity. This makes these different critical behavior easier to distinguish in numerical simulations. The paper is organized as follows: In Section~2 we study the fractal properties of a $c=-5$ conformal field theory coupled to two-dimensional gravity. We define the model in Section~2.1 and discuss the details of the simulations in Section~2.2. In Sections~2.3 and 2.4 we present our measurements of the string susceptibility exponent $\\gamma_s$ and of the fractal dimension $d_H$. And in Section~2.5 we comment on how this particular ensemble of graphs differs from other types of graphs frequently used in studying disordered system. The second part of the paper deals with an Ising model on the $c=-5$ graphs in a quenched approach. In Section~3.1 we discuss the prediction from Liouville theory for the critical behavior of an Ising model coupled dynamically to two-dimensional gravity. In Section~3.2 we discuss details of the simulations and the observables we use to probe the critical behavior. In Sections~3.3 and 3.4 we determine the critical temperature of the Ising model and the corresponding critical exponents. Finally, in Section~4 we summarize and discuss our results. ", "conclusions": "The main results of the work presented in this paper can be summarized as follows: \\begin{itemize} \\item[({\\it a})] The fractal dimension of surfaces, defined by a conformal field theory with central charge $c=-5$ coupled to two-dimensional quantum gravity, is $d_H = 3.36(4)$. This is in reasonable agreement with, and supports, the theoretical prediction Eq.~(\\ref{andH2}), whereas it definitely rules out Eq.~(\\ref{andH1}). \\item[({\\it b})] The critical behavior of an Ising model on a {\\it quenched} ensemble of $c=-5$ graphs agrees well with the predictions, from the KPZ scaling relation, for an Ising model on an {\\it annealed} ensemble of graphs with {\\it identical} fractal properties. \\end{itemize} The first result, especially combined with the recent simulations of $2d$--gravity for $c=-2$ \\cite{cm2}, lends a strong support to Eq.~(\\ref{andH2}) as a correct description of the fractal structure of two-dimensional quantum gravity for $c \\leq 0$. This makes, however, its disagreement with numerical simulations in the region $0 < c \\leq 1$ all the more surprising. What is it in derivation of Eq.~(\\ref{andH2}) that breaks down for $c>0$? Or are the simulations dominated by finite-size errors and simulations of larger systems will eventually agree with Eq.~(\\ref{andH2})? The result for the Ising model is even more interesting. As the theoretical predictions are obtained for an Ising model coupled dynamically to the disorder, this supports the conjecture put forward in the Introduction about the equivalence between annealed and quenched averages over disorder. That is, the only thing relevant for the critical behavior of the Ising model are the average fractal properties of graphs the spins ``see''. How the statistical average over graphs is performed, quenched or annealed, is not relevant. It is also worth noting that we can continuously change the average fractal properties of the graphs by changing the embedding dimension $D$ in Eq.~(\\ref{eq213}). This allows a continuous interpolation between a flat surface and surfaces corresponding to pure gravity. If the prediction of Liouville theory, the KPZ formula, holds for all those models, this implies that the critical behavior of the Ising model should change continuously in the process. In the language of the renormalization group this implies a continuous line of fixed points, rather than isolated points. There are well known examples of this; the low-temperature phase of the two-dimensional $XY$--model or the critical line of the Ashkin-Teller model. But is this statement also true for very weak disorder? If we change the fractal dimension infinitesimally, from 2 to $2+\\epsilon$, is that enough to change the critical behavior of the Ising model? Or, alternatively, does there exist some central charge $c^{\\prime} < -5$ were the geometrical disorder is not strong enough and we always get the Onsager exponents? This point deserves further study. One could also look at the examples of weak disorder that have been studied recently, for example the site or bond-diluted Ising model, and ask if that kind of disorder can also be classified according to some average fractal properties of the lattices. And, moreover, if one could observe some kind of universality in the critical behavior, depending on the fractal structure, akin to what we have presented in this paper. In view of how dramatically the critical behavior of the Ising model changes on surfaces with such strong disorder, one might ask if such change could be observed in real physical systems. Possible candidates for such systems could be, for example, electrons trapped on the interfaces between two liquids, or on the surface of some porous material, were the surfaces had some well defined non-trivial fractal structure. As our results indicate, it is only the average geometry of the surfaces that is important for the Ising model, not its fluid nature or curvature fluctuations. Thus the relative time-scale between the interactions of the particles and the change in the geometry should be irrelevant. \\vspace{20pt} \\noindent {\\bf Acknowledgments:} The work of G.T.\\ was supported by the Humboldt Foundation. The work of P.B.\\ was partially supported by KBN grants 2P03 B19609 and 2P03 B04412." }, "9804/astro-ph9804100_arXiv.txt": { "abstract": "I describe a general framework that could allow to understand the broad band spectra of blazars and lead to a unified picture of the emission from relativistic jets, in BL Lac objects as well as in flat spectrum, radio loud Quasars. The scheme serves as a useful basis to introduce and discuss some of the most interesting results so far obtained on Blazars with {\\it Beppo}SAX. ", "introduction": "The \"blazar phenomenon\" is due to the presence of relativistic flows (jets) emanating from the nuclei of active galaxies which are radio loud. The power to energize the radio lobes is transported in the jets. Blazars are the subset of radio loud AGN for which the relativistic jet happens to point at small angles to the line of sight. Since the radiation emitted by the jet is relativistically beamed into a narrow cone along the direction of motion, the aligned observer will receive a strongly enhanced flux. For a bulk Lorentz factor $\\Gamma\\simeq 10$ at an angle $\\theta\\simeq 1/\\Gamma$ the flux enhancement factor is $10^3-10^4$. The evidence in favor of this picture has been accumulating and is now solid (e.g. \\cite{pu} and refs therein) although the origin of the jets is poorly understood and their physical parameters are highly uncertain. The relativistically amplified, non thermal emission from high energy particles in the jet can account for the extreme properties of blazars concerning variability, polarization and energy distribution of the continuum, which extends from the radio to the gamma-ray band. Traditionally BL Lac objects, where no prominent emission lines are observed (with an upper limit of 5 \\AA), were thought to represent a separate class perhaps more extreme than Quasar-like blazars. The latter include optically violently variable and highly polarized quasars (OVV, HPQ) or more generally Quasars with flat radio spectrum (FSRQ) indicating strong emission from the self absorbed core. It has become clear however that BL Lacs have on average lower luminosity than quasar like blazars (\\cite{pado}) and that the distribution of emission line equivalent widths is continuous (\\cite{scafal}). We will therefore in the following consider blazars as a single class of objects, implicitly assuming that, irrespective of the emission line properties which derive from the surrounding gas, the same physical mechanisms operate in relativistic jets over a wide range of luminosities. By studying the blazar continuum we expect to learn about the radiation mechanisms in the jets, about the processes of particle acceleration and energy transport along the jets and ultimately about their origin and evolution. \\begin{figure*}[bt] \\vspace{9pt} \\psfig{file=sed_medie.ps,width=15.0truecm,height=11.5truecm,rheight=8.7truecm} \\caption{\\small\\sf Average SEDs for the ``total blazar sample'' binned according to radio luminosity irrespective of the original classification. The overlayed dashed curves are analytic approximations obtained assuming that the ratio of the peak frequencies is constant and that the luminosity at the second peak is proportional to the radio luminosity (from \\protect\\cite{foss98} } \\label{fig:sed_medie} \\end{figure*} \\section {The broad band spectra of blazars} The discovery by EGRET on board CGRO of copious $\\gamma$-ray emission from blazars caused a \"Renaissance\" in this field. In fact it had been noted early on that the high density of relativistic electrons necessary to produce the observed compact synchrotron emission would lead to strong, even catastrophic inverse Compton radiation (\\cite{hbs}). X-ray measurements were used to constrain the amount of inverse Compton emission allowed and to derive minimum values for the necessary beaming factors (e.g. \\cite{gp93}). At present the $\\gamma$-ray observations allow to measure the intensity and spectral shape of a component which contains a substantial fraction and in some cases the bulk of the emitted power leading to strong constraints on the physical parameters of the emitting region. It is clearly important, besides observing single objects, to try to derive general properties of the continuum and understand whether and how they differ for instance in BL Lac objects and FSRQ. It is especially interesting to discuss whether the gamma-ray emission is a general property of the whole class. We have recently addressed this problem (\\cite{foss98}) collecting multifrequency data for three complete samples of blazars: the 2 Jy sample of FSRQs, the 1Jy sample of BL Lac objects and the sample of BL Lacs selected in the X-ray band from the Einstein Slew Survey. Systematic differences in the shape of the continuum in specific spectral bands among different subclasses of blazars were noted early on (e.g. \\cite{gg86,imp,ww,smu,umu}. In particular we note that the percentage of objects detected with EGRET (100 MeV - 10 GeV) is significantly larger for the sample of FSRQ than for the two BL Lac samples (40 \\% vs. 26\\% and 17\\% respectively. Nevertheless plotting the average SEDs as shown in Fig.~\\ref{fig:sed_medie}, we can see that the shapes are \"globally similar\". In Fig.~\\ref{fig:sed_medie} all blazars in the three complete samples were merged and grouped in luminosity classes irrespective of their original classification and the dashed lines drawn for comparison derive from an analytic parametric representation (\\cite{foss98}). The main results of this work are the following: \\begin{itemize} \\item two peaks are present in all the SEDs \\item the first peak occurs at lower frequencies for the highest luminosity objects \\item the frequency at which the second peak occurs correlates with that of the first one. The dashed curves correspond to a constant ratio between the two peak frequencies. \\end{itemize} For the most luminous objects the first peak is at frequencies lower than the optical band while for the least luminous ones the reverse is true. Thus highly luminous objects have a \"red\" (steep) IR to UV continuum while objects of lower luminosity have a bluer IR to UV continuum. For this reason and to recall intuitively the location of the peaks on the frequency axis we will briefly call \"red\" blazars the objects in the three highest luminosity classes and \"blue\" blazars those in the two lower luminosity classes. The present data suggest a continuous spectral sequence and no absolute separation between red and blue blazars. Considering the continuum of different objects in a fixed spectral range its shape changes systematically with luminosity along the sequence, as the peak frequency approaches and moves across the chosen frequency interval. In particular the X-ray spectrum becomes steeper and the gamma-ray spectrum (in the EGRET range) becomes flatter from \"red\" to \"blue\" blazars, as the two peaks march to higher frequencies. The different location of the gamma-ray peak can account for the different detection rates of BL Lacs and FSRQs by EGRET. Objects whose $\\gamma$-ray emission peaks in the EGRET range are more easily detected. Recently ground based observations in the TeV range performed with Cherenkov telescopes have detected two of the X-ray brightest \"blue\" blazars (refs). We expect that with the progress in sensitivity many more will be detected giving access to the study of the highest energies from ground. A final comment concerns variability. It is interesting to note that the largest variability is usually observed close to or above each of the two peaks and is usually in the sense of a hardening of the spectrum at higher intensity. These statements are based mostly on observations at high energies (X-rays and gamma-rays) and concern a limited number of sources (e.g. \\cite{umu}) therefore they should be considered as tentative suggestions rather than established facts. \\section {Interpretation} It is generally thought that the first spectral component peaking at far infrared up to X-ray frequencies is due to synchrotron radiation. The spectra from the radio to the submm range most likely involve contributions from different regions of the jet with different self absorption turnovers. However, from infrared frequencies upwards the synchrotron emission should be thin and could be produced essentially in a single homogeneous region. Inverse Compton scattering of soft photons by the high energy electrons emitting the thin synchrotron radiation could be responsible for the second ( high frequency) component of the SED, peaking in the gamma-ray band. The soft photons could be the synchrotron photons themselves (SSC) or photons outside the jet (EC), possibly produced by an accretion disk or torus and scattered or reprocessed by the surrounding gas (e.g \\cite{sbr}, \\cite{umu} and refs therein). If the same region is responsible for the two spectral components then, irrespective of the nature of the seed photons, {\\it the two peaks must derive from the same high energy electrons}. Therefore a change in the density and/or spectrum of those electrons is expected to cause correlated variability at frequencies close to the two peaks. In the SSC model the inverse Compton intensity is expected to vary more than the synchrotron one, approximately as the square of it in the simplest case while in the EC model one expects a linear relation (\\cite{miami}). Measuring the two peaks simultaneously is thus the best means to determine the physical parameters of the emission region and studying the variability of the spectra around the peaks yields unique insight into the mechanisms of particle acceleration and energy loss in the jet. The variability correlation should enable to disentangle the contribution of different sources of seed photons (SSC vs. EC). The \"spectral sequence\" discussed above could be attributed to a systematic dependence of the critical electron energy (the break energy) and/or of the magnetic field on luminosity. Assuming that the beaming factors are not significantly different along the sequence, the trend in apparent luminosity is also a trend in intrinsic luminosity. In the SSC model the break energy of the electrons is univocally determined by the ratio of the frequencies of the two peaks and should therefore be approximately constant. \"Red\" blazars should then have lower magnetic field than \"blue\" blazars. Systematic model fitting of all the $\\gamma$-ray detected blazars with sufficient multifrequency data suggest that as the magnetic energy density decreases the external photon energy density becomes important so that a smooth transition between the SSC and EC scenario takes place (\\cite{gg98}). \\section {SAX observations} The X-ray band is crucial for a discussion of the above problems in that the synchrotron and inverse Compton components which have different spectral shapes may both be relevant. Simultaneous observations over a broad energy range are required to disentangle the two mechanisms. The unique characteristics of the {\\it Beppo}SAX instrumentation appear therefore ideal for blazar studies. Observations of bright blazars detected in $\\gamma$--rays were proposed with the main aims of: \\begin{itemize} \\item determining the spectral shape up to the 100 keV range, thus exploring the connection between X-rays and gamma-rays \\item studying the variability in relation with other wavebands, especially $\\gamma$--rays. \\end{itemize} \\begin{figure*}[bt] \\vspace{9pt} \\psfig{file=sed_3c279.ps,width=15.0truecm,height=13truecm,rheight=11.2truecm} \\caption{\\small\\sf Spectral energy distribution during the 1997 campaign, compared with previous ones. The 1997 data are from: X--rays = {\\it Beppo}SAX data; $\\gamma$--rays = Hartman, private comm.; R-band = Raiteri \\& Villata, private comm.} \\label{fig:sed_3c279} \\end{figure*} In the following I will briefly mention and comment some of the most interesting results obtained so far. Besides 3C 273 (\\cite{htv}) which is probably intermediate between a blazar and a \"normal\" quasar, two flat spectrum radio quasars, 3C 279 and PKS 0528+134, were observed (before June 1997) and found in a low intensity state. In the scheme presented above these are \"red\" blazars. I will discuss here the first source (also \\cite{mtv}), while for the second one I refer to \\cite{geltv}. Finally I will consider results on \"blue\" blazars ( Mrk 421, 1ES 2344+514 and Mrk 501 (see the contributions in this volume \\cite{ftv}, \\cite{gitv} and \\cite{gtv} respectively). It is important to remember that in \"red\" blazars the X-ray emission represents the lower energy end of the inverse Compton emission, while for \"blue\" blazars it is the high energy end of the synchrotron emission. \\subsection {3C 279} The X-ray spectrum measured with {\\it Beppo}SAX in January 1997, the simultaneously measured optical flux and the quasi simultaneous gamma-ray flux (Hartman, private communication) are shown in Fig.~\\ref{fig:sed_3c279} together with other simultaneously measured SEDs obtained at other epochs: the high state observed in June 1991 (\\cite{hart}), the low state observed in January 1993 (\\cite{m94}) and the preflare and flare states observed in January - February 1996 (\\cite{wehrle98}). At the epoch of the {\\it Beppo}SAX observations the $\\gamma$--ray flux was a factor 6 and 20 weaker respectively than measured in June 1991 and early February 1996. In X-rays the amplitude is smaller but there is a good correlation between the X-ray and gamma-ray fluxes especially in the 2-10 keV band (see Fig. 3 in Maraschi et al. this volume). Note that the fluxes at 1 keV measured by {\\it ROSAT} and {\\it Beppo}SAX in the 1993 and 1997 low states are similar However the spectrum measured by {\\it Beppo}SAX is significantly flatter providing a good connection with the higher $\\gamma$--ray flux in 1997. The simultaneity (within one day) of the X-ray (XTE) and $\\gamma$--ray peaks during the 1996 flare suggests that the X-ray to gamma-ray emission originates in a single region and that the spectrum hardens with increasing intensity. It is possible that the $\\gamma$-ray peak in the SED moves to higher energies at the flare peak. The situation is much more complex at lower energies. Although there is still a general correlation of the IR-optical-UV fluxes with the gamma-ray intensity on long timescales, the flux variation at optical wavelengths corresponding to the rapid 1996 flare is quite small. Note also that for 3C 279 the (presumed) peak of the synchrotron component falls in an unexplored region of the spectrum, between $10^{12}$ and $10^{14}$ Hz. In the SSC model one expects that the inverse Compton emission varies with the square of the amplitude of the synchrotron emission due to the same electrons (e.g. at the two peaks of the SED). This is compatible with the long term variations but not with the strong rapid flare observed in 1996, where the amplitude in gamma-rays was larger than the square of the optical one. On the other hand if the seed photons for the inverse Compton process are external to the jet, they should not be rapidly variable and the inverse Compton emission is expected to vary linearly with the synchrotron one. Thus neither of the two \"simple\" models can adequately account for the multifrequency variability behaviour. A possible way out is that the seed photons derive from backscattering and/or reprocessing of radiation produced in the jet by gas clouds closely approaching the jet itself (\\cite{ggmadau}). This model is attractive and needs to be studied in more detail. Another possible way out is that the region emitting the synchrotron radiation is inhomogeneous so that the observed variability is diluted by a more stationary component. \\subsection {Blue blazars} The X-ray emission from these objects has been observed to vary dramatically on short timescales at least in the brightest prototypes, PKS 2155-304 and Mrk 421. This can be understood recalling that the X-ray emission represents the high energy end of the synchrotron component: it is therefore due to radiation from the highest energy electrons which have the shortest lifetimes and can vary very rapidly. A continuous acceleration mechanism must be responsible for maintaining in the source particles which have lifetimes of the order of hours. However the injection mechanism may be continuous only in an average sense or at a low intensity level and episodes of increased injection rate may occur causing variability. The study of flares and of spectral variability associated with them gives direct information on the spectra of the freshly injected/accelerated particles and their subsequent decay to a state of quasi-equilibrium. The photons upscattered through the inverse Compton process by the highest energy electrons reach TeV energies so that the \"bluest\" and brightest sources can be detected from ground based Cherenkov telescopes. This is the case up to now for three objects: Mrk 421, 1ES 2344+514 and Mrk 501. All of them were observed with SAX in the first half of AO1. The simultaneous observation of a source in X-rays and at TeV energies should allow to determine unambiguously the energy of the radiating electrons, the beaming factor, the magnetic field and the energy density of the seed photons. In some cases it may be necessary to take into account that scattering will occur in the Klein Nishina regime. \\begin{figure*}[bt] \\vspace{9pt} \\psfig{file=sed_mkn421.ps,width=15.0truecm,angle=270,height=13truecm,rheight=11.2truecm} \\caption{\\small\\sf The May 97 data from {\\it Beppo}SAX are compared with the energy distributions measured in 1994 (\\protect\\cite{macomb}} \\label{fig:sed_mkn421} \\end{figure*} \\subsection {Mrk 421} This source has been repeatedly observed with ASCA and at other relevant wavelengths including TeV observations. (\\cite{macomb}) The {\\it Beppo}SAX observations of Mrk 421 (Fossati et al. this volume) show a decay between two intensity states closely similar to those previously observed with ASCA. The higher state shows a flatter spectrum indicating that the emission peaks at higher energies. In Fig.~\\ref{fig:sed_mkn421} the {\\it Beppo}SAX data are compared with the ASCA and TeV data presented by \\cite{macomb}. Mastichiadis and Kirk \\cite{mk}have shown that this behaviour could be the result of an injection mechanism in which the maximum energy of the injected particles increases. The two \"states\" would therefore represent equilibrium states for injection spectra extending to different maximum energies. The fact that the spectral variability observed in 1994 is closely reproduced in the {\\it Beppo}SAX observations indicates that the involved region is the same (same physical parameters). In addition SAX detected the source at higher energy with the PDS. The preliminary analysis yields a flat spectrum at high energies suggesting that the inverse Compton component becomes dominant in this band. The variation in the PDS is much smaller than in the MECS which is consistent with attributing the hard emission to IC from electrons of much lower energies. An extensive campaign for simultaneous ASCA and TeV observations will take place in the spring of 1998, to which {\\it Beppo}SAX observations could add significantly. \\subsection {1ES 2344+514, Mrk 501} This source is not as well studied as Mrk 421 but shows a similar, more extreme behaviour (\\cite{gitv}). The emission peak was in the medium X-ray range in December 96 and shifted to 20-50 keV during the flare of a factor 2 observed between December 3-7 96. The most extraordinary spectral variation was found in Mrk 501 (\\cite{pian98}, \\cite{gtv}). The source was in a state of strong activity at TeV energies although the 1 keV flux had not increased dramatically. However compared to previous observations {\\it the X-ray spectrum had changed dramatically} being flatter than or close to 1 up to 100 keV. In correspondence to a large TeV flare , the 1 to 100 keV spectrum hardened still indicating that the emission peak was at or above 100 keV, while past multifrequency measurements showed it to be below 1 keV. Thus in Mrk 501 the shift of the peak frequency was more than a factor 100. Some theoretical implications are discussed by Ghisellini (\\cite{gtv}). This unprecedented behaviour may be less uncommon than judged at first sight. In fact in the medium X-ray band the intensity behaviour was not exceedingly dramatic and good spectral capabilities up to the hard X-ray band, necessary to reveal the phenomenon, have only recently become available with {\\it Beppo}SAX. It is interesting to mention that a spectral survey of BL Lacs selected in relatively soft X-rays yields evidence of some hard X-ray spectra (\\cite{wolter}). Sources which have more or less permanent emission peaks in the hard X-ray range may also exist and have gone undetected so far. ", "conclusions": "The results from the first part of the {\\it Beppo}SAX AO1 Core Program on Blazars are extremely exciting. Bright $\\gamma$-ray blazars can be usually detected up to 50 - 100 keV allowing detailed study of the synchrotron and inverse Compton emission and of the correlations between their temporal and spectral variability. The observations presented above suggest and support: \\begin{itemize} \\item the correlation between the X--ray (1 keV) and $\\gamma$--ray (0.1-10 GeV) fluxes, which now holds over a period of 6 years and about a factor 30 in flux change for $\\gamma$-rays (3C 279 -- PKS 0528 + 154) \\item the smaller amplitude of variability of the synchrotron component compared to the high energy one. \\item the synchrotron emission peak seems to shift systematically to higher energies during flares (Mrk 421, 1ES 2344+541, Mrk 501) \\end{itemize} These results will undoubtedly stimulate new observations and new theoretical approaches. In particular genuinely time dependent models for the acceleration of particles and the spectral evolution of the emitted radiation are needed." }, "9804/astro-ph9804336_arXiv.txt": { "abstract": "In this paper we present a detailed study of the radio galaxy J1324$-$3138, located at a projected distance of 2$^{\\prime}$ from the centre of the Abell cluster of galaxies A3556, belonging to the core of the Shapley Concentration, at an average redshift z=0.05. We have observed J1324$-$3138 over a wide range of frequencies: at 327 MHz (VLA), at 843 MHz (MOST), and at 1376 MHz, 2382 MHz, 4790 MHz and 8640 MHz (ATCA). Our analysis suggests that J1324$-$3138 is a remnant of a tailed radio galaxy, in which the nuclear engine has switched off and the radio source is now at a late stage of its evolution, confined by the intracluster gas. The radio galaxy is not in pressure equilibrium with the external medium, as it is often found for extended radio sources in clusters of galaxies. We favour the hypothesis that the lack of observed polarised radio emission in the source is due to Faraday rotation by a foreground screen, i.e. the source is seen through a dense cluster gas, characterised by a random magnetic field. An implication of the head-tail nature of the source is that J1324$-$3138 is moving away from the core of A3556 and that possibly a major merging event between the core of A3556 and the subgroup hosting J1324$-$3138 has already taken place. ", "introduction": "Extended radio emission associated with galaxies in clusters is often characterised by morphologies which reflect the interaction between the radio emitting plasma and the local environment in the cluster. Head-tail sources are usually associated with non-dominant cluster galaxies moving at a considerable speed within the cluster. Their morphologies are then explained as the result of ram pressure exerted by the intergalactic medium on the double sided radio emission (see for example O'Dea \\& Owen 1985a and 1985b, and Owen 1996, for a recent review). Wide-angle tail radio galaxies, on the other hand, are more difficult to account for with the above mentioned model, since they are usually associated with massive and dominant cluster galaxies, with much lower peculiar velocities with respect to the cluster mean. Beyond ram pressure, it is now accepted that large flows of hot gas could provide a wind within clusters of galaxies able to bend straight jets into wide-angle tail morphologies (Owen 1996 and references therein). \\noindent The study of extended galaxies in clusters is important for a variety of reasons. The morphology and the direction of the extension may give important information on the dynamics of the galaxy, such as, for example, the direction of the motion projected on the plane of the sky. Furthermore, the non-thermal pressure in the tails of radio emission can be compared to the thermal pressure exerted by the intracluster gas, in those cases where X-ray data are available to provide estimates of the temperature and pressure. This is crucial for studying the interaction between the radio emission and the external gas, and for deriving information on the evolution of radio sources in clusters of galaxies, as well as the influence of the cluster dynamics (such as, for example, merging processes) on the radio properties of the cluster. Last but not least, the observed polarisation properties of the radio emission may give information on the intracluster magnetic field and its structure. \\medskip In this paper we present a detailed study of the extended radio galaxy J1324$-$3138 (RA$_{J2000} = 13^h24^m01^s$, DEC$_{J2000} = -31^{\\circ}38^{\\prime}$), located in the central region of the Abell cluster A3556. It was first observed in a radio survey of the clusters of galaxies in the Shapley Concentration core carried out at 843 MHz with the Molonglo Observatory Synthesis Telescope (MOST) and at 1376 MHz with the Australia Telescope Compact Array (ATCA) (Venturi et al. 1997, hereinafter Paper I). This work is part of a larger project whose aim is to study the radio/optical properties of the clusters in the core of the Shapley Concentration, in particular the chain formed by A3556-A3558-A3562 (Venturi et al. 1998), both from a statistical point of view and through a detailed analysis of the physical properties of the extended radio galaxies in these clusters. In Figure 1 the superposition of the radio isophotes on the Digitised Sky Survey shows that the radio component located at the south west extremity of the extended radio emission is coincident with the nucleus of the 15.6 magnitude galaxy \\#5975 in the COSMOS catalogue (RA$_{J2000}$ = $13^h23^m57.5^s$, DEC$_{J2000}$ = $-31^{\\circ}38^{\\prime}45^{\\prime\\prime}$). Its radial velocity velocity v = 15054 km s$^{-1}$ (Stein 1996) establishes that it belongs to A3556 ($ = 14357$ km s$^{-1}$, Bardelli et al. 1998). This, coupled with the fact that only a few very faint optical objects fall within the envelope of the radio emission, led us to the conclusion that J1324$-$3138 is an extended, possibly head-tail, radio galaxy located in the vicinity of the cluster centre (see Paper I). In Section 2 we present the observational data. In Section 3 the morphology of the source is described and analysed, and in Section 4 a detailed study of the synchrotron spectrum is carried out. The nature of the source, its relation to the cluster of galaxies A3556 and its implications for cluster merging and formation is discussed in Section 5. Throughout the paper we use a Hubble constant of H$_0$ = 100 km s$^{-1}$Mpc$^{-1}$. At the redshift of the cluster this implies that 1$^{\\prime\\prime}$ = 0.67 kpc. \\begin{figure} \\epsfysize=8.5cm \\epsfbox{FIG1.PS} \\caption[]{ 4790 MHZ radio isophotes of the extended radio galaxy J1324$-$3138 and of the nearby radio galaxy J1324$-$3140, associated with the dominant cD galaxy in the centre of A3556 (see Paper I), superimposed on the DSS optical image. The resolution of the image is $20^{\\prime\\prime} \\times 10^{\\prime\\prime}$, p.a. $0^{\\circ}$.} \\end{figure} \\vskip 1.0truecm \\noindent ", "conclusions": "We have presented observations of the radio galaxy J1324$-$3138, located in the central region of the Abell cluster A3556, over a wide range of frequencies and resolutions. We can briefly summarise our results as follows: \\noindent {\\it (a)} J1324$-$3138 is an example of a {\\it remnant} of a radio galaxy, i.e. a source in which the engine of the radio emission has switched off. The evolution of the radio emission is presently dominated by synchrotron losses. \\noindent {\\it (b)} The radio source is not in pressure equilibrium with the intracluster gas. In particular it is underpressured. \\noindent {\\it (c)} We suggest that the lack of polarisation in the source is due to the presence of an intervening Faraday screen, i.e. a cluster scale magnetised medium, as it is now often observed in clusters of galaxies, which depolarises the radio emission. \\noindent {\\it (d)} Under the hypotesis that cluster mergers influence the radio emission of a galaxy, the properties of J1324$-$3138, coupled with the peculiarities of A3556 at radio and optical wavelengths (Paper I and Bardelli et al. 1998), suggest that the core of A3556 and the subgroup hosting J1324$-$3138 have already undergone a major merging event. \\vskip 1.0truecm \\noindent {\\bf Acknowledgments} We wish to thank D. Dallacasa for his suggestions and discussion while this work was carried out, and R. Fanti for careful reading of the manuscript. We are grateful to S. Ettori for providing unpublished results, and to P. Stein for giving us the spectrum of J1324$-$3138. T.V. acknowledges the receipt of two grants from CNR/CSIRO (Prot. n. 119816 and Prot. n. 088864). The MOST is operated by the University of Sydney, with support from the Australian Research Council. The Australia Telescope Compact Array is operated by the CSIRO Australia Telescope National Facility. The National Radio Astronomy Observatory (NRAO) is operated by Associated Universities, Inc., under contract with the National Science Foundation. This research has made use of the NASA/IPAC Extragalactic Database (NED), which is operated by the Jet Propulsion Laboratory, Caltech, under contract with the National Aeronautics and Space Administration." }, "9804/astro-ph9804046_arXiv.txt": { "abstract": "The CAT (Cherenkov Array at Th\\'emis) imaging telescope, equipped with a very-high-definition camera (546 fast phototubes with $0.12^{\\circ}$ spacing surrounded by 54 larger tubes in two guard rings) started operation in Autumn 1996 on the site of the former solar plant Th\\'emis (France). Using the atmospheric Cherenkov technique, it detects and identifies very high energy $\\gamma$-rays in the range $250\\:{\\mathrm GeV}$ to a few tens of TeV. The instrument, which has detected three sources (Crab nebula, Markarian 421 and Markarian 501), is described in detail. ", "introduction": " ", "conclusions": "" }, "9804/astro-ph9804270_arXiv.txt": { "abstract": "We report results of a \\rosat\\ High-Resolution Imager (HRI) observation of the X-ray error box given by the \\sax\\ Wide Field Camera for the gamma-ray burst that occurred on 1997 February 28. The observation started 10 days after the burst and ended three days later, with a total exposure of 34.3~ks. An X-ray source was detected within the 3$'$ WFC error box and its position determined with a 10$''$ radius accuracy. The source position is in the \\sax\\ Narrow Field Instrument source error box and is coincident (to within 2$''$) with the optical transient associated with GRB970228. This is the most precise position obtained for an X-ray afterglow and confirms that the X-ray and optical afterglows have the same origin. We present the 0.1--2.4~keV combined HRI and \\sax\\ Low-Energy Concentrator Spectrometer decay light curve which can be well fit with a power-law. The decay is consistent with that measured at higher energies (2--10~keV) with the \\sax\\ Medium-Energy Concentrator Spectrometer. ", "introduction": "Observations of celestial Gamma-Ray Bursts (GRB) performed over the last 25 years had not, until recently, succeeded in finding counterparts in other wavelength bands. The ability of the \\sax\\ satellite to provide arc minute precision positions (Piro et al. 1998) and to observe these positions within hours of the GRB changed this situation in 1997 when the X-ray afterglow of GRB970228 was measured (Costa et al 1997a). The burst was detected (Costa et al. 1997a) with the Gamma-Ray Burst Monitor (GRBM) (40--70~keV, Frontera et al. 1997a) on 1997 February 28.123620 UT and also detected in the 1.5--26 keV energy range by one of the two Wide Field Cameras (WFC No. 1) aboard the same satellite (Jager et al. 1997). Its position was determined with an error circle of 3~arcmin (3$\\sigma$) radius, centered on $\\alpha_{2000}\\,=\\,05^h01^m57^s$, $\\delta_{2000}\\,=\\,11^\\circ46'24''$. Eight hours after the GRB trigger, from February 28.4681 to February 28.8330 UTC, the Narrow Field Instruments (NFI) on board \\sax\\ (Boella et al. 1997a) were pointed to the WFC error box. An X-ray source, SAX J0501.7+1146, was detected (Costa et al 1997b) in the field of view of both the Low Energy (0.1--10~keV) and Medium Energy (2--10~keV) Concentrators Spectrometers (LECS and MECS) (Parmar et al. 1997; Boella et al. 1997b). The source position ($\\alpha_{2000}\\,=\\,05^h01^m44^s$, $\\delta_{2000}\\,=\\,11^\\circ46'42''$) is consistent with the GRB error circle. The source was again observed about three days later, from March 3.7345 to March 4.1174. During this observation, the 2--10~keV source flux had decreased by about a factor 20, while in the 0.1--2~keV energy range the source was not detected. Following the discovery of the GRB, searches for radio and optical counterparts to GRB970228 were conducted with most of the ground based telescopes in the northern hemisphere. Groot et al. (1997) reported the discovery of an optical transient at a position ($\\alpha_{2000}\\,=\\,05^h01^m46.70^s$, $\\delta_{2000}\\,=\\,11^\\circ46'53.0''$), consistent with both the \\sax\\ WFC and NFI error boxes and with the long baseline timing \\ulysses/\\sax\\ and \\ulysses/\\wind\\ error annuli, of 31$''$ and 30$''$ half-width, respectively (Hurley et al. 1997; Cline et al. 1997). While the association of the transient X-ray source with the afterglow of GRB970228 was compelling on the basis of the properties of its decay curve when extrapolated backwards to the burst time (Costa et al. 1997c), the association of the optical transient with the burst afterglow was less strong. In spite of the positional consistency and temporal behaviour of the optical transient, it was not possible to exclude the possibility that the optical transient was unrelated to the GRB (see discussion by van Paradijs et al. 1997), like in the case of the radio source discovered in the earliest error box of GRB970111, which showed a time behaviour consistent with that expected from radio afterglows of GRBs, but later resulted to be unrelated tho the burst (Frail et al. 1997). The \\rosat\\ satellite, thanks to its HRI focal plane instrument, offered the possibility of imaging the X-ray afterglow at 10$''$ angular resolution (David et al. 1997). A Target of Opportunity observation was thus requested and obtained. Here we report on results of that observation and its consequences. Preliminary results were already previously reported (Frontera et al. 1997b). ", "conclusions": "The \\rosat\\ HRI observation of the \\sax\\ WFC error circle of GRB970228 clearly shows the presence of a new X-ray source. Its position within the error box of the \\sax\\ source, the low probability of a chance coincidence ($\\sim 1 \\times 10^{-3}$) and the better imaging capabilities of the \\rosat\\ HRI compared to the \\sax\\ NFI, indicate that the \\rosat\\ source and the \\sax\\ source are the same object. The source position derived from the \\rosat\\ observation is the most precise position of a GRB X-ray afterglow obtained thus far. Its position is also coincident with the optical transient associated to GRB970228 within 2$''$. This result confirms that X-ray source and the optical transient are the same object. The X-ray source hows a 0.1--2.4 keV decline according to a power law decline with index $\\alpha \\,=\\, 1.50^{+0.23}_{-0.35}$ for at least 13 days. This slope is fully consistent with that estimated in the 2--10 keV energy band (Costa et al. 1997c) and is marginally consistent with that reported by Fruchter et al. (1997, 1998) in the optical band. Thus it appears that from the X-ray to the optical band the GRB afterglow has the same decline law. \\begin{figure} \\epsfig{figure=grb282_l_tot.ps,height=8.5cm,width=8.5cm,angle=0} \\caption{As in Fig. 2, but extrapolated to the first second from the burst onset. The two arrows on the left delimit the time interval, that corresponds to the GRB last three pulses, when the X-ray afterglow is expected to start (see text). } \\label{figure:decayb} \\end{figure}" }, "9804/astro-ph9804264_arXiv.txt": { "abstract": "Including nucleon--nucleon correlations due to both Fermi statistics and nuclear forces, we have developed a general formalism for calculating the charged--current neutrino--nucleon absorption rates in nuclear matter. We find that at one half nuclear density many--body effects alone suppress the rates by a factor of two and that the suppression factors increase to $\\sim$5 at $4\\times10^{14}$ g cm$^{-3}$. The associated increase in the neutrino--matter mean--free--paths parallels that found for neutral--current interactions and opens up interesting possibilities in the context of the delayed supernova mechanism and protoneutron star cooling. ", "introduction": "The neutrino absorption and scattering opacities in the post--shock core of a supernova, in which nuclei are largely disintegrated into nucleons, determine the duration, spectrum, and flavor distribution of the emerging neutrino pulse. It has been known for some time that the interactions among nucleons in the denser regions can change these opacities significantly, but to date there has been no comprehensive treatment given in the literature and present calculations of the complete supernova process do not include the effects of interactions on the opacities. The neutrino--matter interaction rates can be related to the space-- and time--dependent correlations among the set of density operators for the separate nuclear constituents (to find the Gamow--Teller parts we must consider separate spin--up and spin--down densities). In the case of neutral--current interactions \\cite{BS}, there is an instructive limit, which also provides an estimate of the effects, in which the combined limits of large nucleon mass and small neutrino energy allow the use of long--wavelength limits of equal--time correlation functions, in turn expressible in terms of the second derivatives of an energy density functional with respect to various densities. This approach is the direct multichannel generalization of the familiar results for light scattering from the thermal density fluctuations in a fluid, where it is the compressibility that determines the long--wavelength opacity, and it was used in references \\cite{s2Ray} and \\cite{iwa} to find significant reductions of neutral--current opacity in certain regions. In Burrows \\& Sawyer \\cite{BS}, an approach based on ring graphs was used to encompass these results and to extend them to domains in which the equal--time and long--wavelength limits are not clearly applicable. The use of the equal--time and long--wavelength limits to express correlation functions in terms of static susceptibilities cannot be extended to the charged--current interactions when there is a large chemical potential difference between protons and neutrons. Furthermore, there do not exist in the present literature systematic estimates of the effects of interactions on the charged--current opacities for electron neutrinos. In the present work, we give a theoretical framework for addressing these opacities, based on summing ring graphs, together with the results of calculations with input parameters taken from the current phenomenology of nuclear matter. ", "conclusions": "We have developed a new formalism for incorporating the effects of many--body correlations on the charged--current rates of neutrino--matter interactions. This formalism reveals that these rates are considerably suppressed in the densest regions of protoneutron stars and supernova cores. Assuming that the nucleons are non--relativistic, our formalism incorporates the full kinematics of the interaction, Pauli blocking by final--state nucleons (protons), and correlation due to nucleon--nucleon interactions. We have employed the ring approximation (RPA) and assumed the near--validity of Fermi Liquid Theory. It would desirable to include ladder diagrams and to perform the calculations in the context of a better numerical method for solution of the nuclear equation of state (EOS), since the solution of the EOS is intimately related to the derivation of the scattering/absorption rates. However, those who perform detailed nuclear EOS calculations and address many--body correlations in nuclear matter do not as yet provide the requisite spin and density structure functions, even for the static case. These results for charged currents, when combined with the results from Burrows \\& Sawyer \\cite{BS} for neutral currents, strongly suggest that energy and lepton number will leak from supernova cores at a rate that is higher than heretofore estimated. This implies that the neutrino luminosities during the epoch after bounce for which the inner core is the major energy source ($> 0.5 - 1.5$s) will be enhanced, perhaps by as much as 50\\% \\cite{BS}. The consequences of this increased transparency for the neutrino--driven supernova explosion mechanism \\cite{bhf} may be interesting, but have yet to be clarified." }, "9804/astro-ph9804052_arXiv.txt": { "abstract": "ACO 3627 is a rich, nearby cluster of galaxies at the core of the Great Attractor. At the low galactic latitude of $b = -7.2\\deg$ the galactic extinction is significant. Nevertheless, its proximity makes it a prime target for studies of environmental effects on its cluster members. Here, we report on a multi-wavelength study of a Seyfert 1 galaxy at 30 arcmin from the centre of ACO 3627. Its Seyfert nature was discovered spectroscopically and confirmed in X-rays. We have obtained B$_{\\rm J}$ and R$_{\\rm C}$ CCD photometry as well as J, H, K and L aperture photometry at the SAAO, low and high resolution spectroscopy (ESO and SAAO), 21 cm line observations (Parkes Observatory) and X-ray ROSAT PSPC data. The Seyfert 1 galaxy is of morphology SBa(r). It has a nearby companion (dS0) but shows no signs of interaction. A consistent value for the galactic extinction of A$_{\\rm B}$ = 1.6 mag could be determined. The nucleus of the Seyfert is very blue with a strong (B$_{\\rm J}$ -- R$_{\\rm C}$) colour gradient in the inner 2.5 arcsec. The extinction-corrected near-infrared colours of WKK 6092 are typical of a Seyfert 1 and the X-ray spectrum conforms to the expectation of a Seyfert as well. The galaxy has a very low \\HI\\ flux. This could be explained by its morphology, but also -- due to its very central position within the rich Norma cluster -- to ram pressure stripping. ", "introduction": "Dust and stars in the Milky Way obscure a large fraction of the extragalactic sky, creating a ``Zone of Avoidance'' (ZOA) in the distribution of galaxies. In an effort to reduce the size of the ZOA and thus coming closer towards an all-sky distribution of galaxies, we have embarked on a deep optical galaxy search behind the southern Milky Way (Kraan-Korteweg \\& Woudt 1994). This has led to the recognition that ACO 3627 (Abell \\etal\\ 1989), also called the Norma cluster after the constellation it is located in, is a massive, nearby cluster of galaxies at the core of the Great Attractor (GA) $(\\ell,b,v) = (325\\deg, -7\\deg, 4882$ \\kms) (Kraan-Korteweg \\etal\\ 1996). The Norma cluster appears to be the central, dominant component of a ``great wall''-like structure and would be the most prominent overdensity of galaxies in the southern sky, were it not obscured by the Milky Way (Woudt \\etal\\ 1997). Recent observations of the Norma cluster with the ROSAT PSPC have confirmed the massive nature of ACO 3627; it is the 6$^{th}$ brightest cluster in the ROSAT sky (B\\\"ohringer \\etal\\ 1996). The X-ray contours furthermore suggest the existence of a subcluster. The merging scenario is independently supported by the radio continuum emission of the central cD galaxy PKS1610-608. The emission from this wide-angle-tail (WAT) radio source (Jones \\& McAdam 1992) seems to encircle the X-ray subcluster (\\cf\\ Fig.~3 of Kraan-Korteweg \\etal\\ 1997) and is indicative of a strong motion of the cluster gas due to the ongoing merging process (Jones \\& McAdam 1996, Burns \\etal\\ 1994). Roughly 30{\\arcmin} from the centre of this cluster -- taken as the central cD galaxy PKS1610-608 -- we have identified a Seyfert 1 galaxy. It is a member of ACO 3627. In the following sections we describe the various observations of this galaxy: the discovery of the galaxy in section 2, the multicolour photometry obtained at the South African Astronomical Observatory (SAAO) in section 3, the spectroscopy obtained at the European Southern Observatories (ESO) and the SAAO in section 4, the \\HI\\ observations obtained with the 64m radio telescope of the Parkes Observatory of the Australian Telescope National Facility (ATNF) in section 5, and the X-ray data from ROSAT PSPC observations in section 6. The results are summarized and discussed in the last section. ", "conclusions": "We have observed the Seyfert galaxy WKK 6092 at different wavelengths. The resulting data are summarized in Table~\\ref{seytab}. \\begin{table} \\caption{Observational parameters of WKK 6092} \\label{seytab} \\begin{tabbing} Apparent blue (IIIaJ)(m$_{B}$) magn \\= \\kill {\\bf Coordinates:} \\> \\\\ R.A. (1950) \\> $16^{h} 07^{m} 32.7^{s}$ \\\\ DEC. (1950) \\> $-60^{\\circ} 30' 11''$ \\\\ Galactic Longitude \\> $325.20^{\\circ}$ \\\\ Galactic Latitude \\> $-6.74^{\\circ}$ \\\\ \\vspace{5mm} {\\bf Properties:} \\> \\\\ Hubble Type \\> SBa(r) \\\\ Dimensions (a x b) \\> 56'' x 47'' \\\\ Ellipticity (1-b/a) \\> 0.11 \\\\ Inclination \\> 28$\\deg$ \\\\ Position Angle \\> 96$\\deg$ \\\\ \\vspace{5mm} {\\bf Photometry:} \\> \\\\ $B_{\\rm J}$ (IIIaJ) \\> 14.7 $\\pm$ 0.5 mag \\\\ $B_{25}$ (CCD) \\> 14.96 $\\pm$ 0.09 mag \\\\ $B_{\\rm T}$ (CCD) \\> 14.88 $\\pm$ 0.13 mag \\\\ $R_{24}$ (CCD) \\> 13.38 $\\pm$ 0.12 mag \\\\ $R_{\\rm T}$ (CCD) \\> 13.30 $\\pm$ 0.14 mag \\\\ J$_{\\rm c}$ \\> 12.91 $\\pm$ 0.03 mag \\\\ H \\> 11.98 $\\pm$ 0.03 mag \\\\ K \\> 11.51 $\\pm$ 0.03 mag \\\\ L \\> 10.60 $\\pm$ 0.20 mag \\\\ \\HI\\ flux \\> 0.93 Jy \\kms \\\\ X-Ray (0.5--2.0 keV): \\> \\\\ \\hspace{0.25cm} Flux \\> $1.05 \\pm 0.15 \\cdot 10^{-12}$ erg s$^{-1}$ \\\\ \\> \\hspace{2.65cm} cm$^{-2}$ \\\\ \\hspace{0.25cm} ${\\cal L}_X$ \\> $1.2 \\pm 0.17 \\cdot 10^{42}$ erg s$^{-1}$\\\\ \\vspace{5mm} {\\bf Galactic Extinction (A$_B$):} \\> \\\\ from HI \\> 1.5 mag \\\\ from Balmer decrement \\> $\\la$ 1.7 mag \\\\ from X-ray \\> 1.6 mag \\\\ \\vspace{5mm} {\\bf Heliocentric velocity:} \\> \\\\ MEFOS \\> 4711 $\\pm$ 30 \\kms \\\\ S.A.A.O. \\> 4688 $\\pm$ 40 \\kms \\\\ Parkes 64-m \\> 5012 $\\pm$ 5 \\kms \\\\ \\hspace{0.25cm} \\DVF \\> 88 \\kms \\\\ \\hspace{0.25cm} \\DVT \\> 97 \\kms \\\\ \\end{tabbing} \\end{table} WKK 6092 and its neighbour are both members of ACO 3627. They have similar redshifts but show no indications of interaction. The morphology of both galaxies do not seem distorted (\\cf\\ Fig.~\\ref{ccdim}). The Seyfert has a very blue nucleus, a distinct bar and a ring superimposed on an otherwise smooth disk. An upper limit for the Galactic foreground extinction in the line of sight of the Seyfert galaxy can be set at A$_{\\rm B}$ = 1.6 mag. This was determined by three different methods, the Balmer decrement in the optical spectrum, the fitting of an absorbed power low to the X-ray spectrum and the Galactic \\HI\\ column densities. All give a consistent value of the foreground extinction. A minor fraction of the extinction is intrinsic to the galaxy. The extinction corrected near-infrared colours of WKK 6092 are typical of a Seyfert 1 and are in agreement with well known Seyfert 1's such as NGC 1566 (Glass and Moorwood 1985). The X-ray sepctrum is also consistent with the standard expectation for this object. At the adopted cluster distance of R = 93 $h_{50}^{-1}$ Mpc, the absolute magnitude (corrected for the galactic extincton) is M$_{B_T}^o = -21.52$. The \\HI\\ and total mass is $1.9 \\cdot 10^{9} {\\cal M}_{\\odot}$ and $30 \\cdot 10^{9} {\\cal M}_{\\odot}$, respectively. The Seyfert is at a projected distance of 0.8 $h_{50}^{-1}$ Mpc from the cluster centre and the \\HI\\ content of the galaxy might be influenced by interactions with the Inter Cluster Medium due to processes like ram pressure stripping (Cayatte \\etal\\ 1990). The \\HI\\ content is in fact quite low. This is, however, not inconsistent with the expectation for a barred early-type spiral. Despite the difficulties in analysing data of an object deep within the Milky Way, all data concerning the here investigated Seyfert galaxy WKK 6092 at 30 arcmin from the centre of the rich cluster ACO 3627 correspond to the standard characteristics of a Seyfert 1 galaxy." }, "9804/astro-ph9804322_arXiv.txt": { "abstract": "We report the {\\em first} unambiguous detection of the host galaxy of a normal radio-quiet QSO at high-redshift in $K$-band. The luminosity of the host comprises about 35\\%\\ of the total $K$-band luminosity. Assuming the average colour of QSOs at $z\\approx2$, the host would be about 5 to 6~mag brighter than an unevolved $L_*$ galaxy placed at $z\\approx2$, and 3 to 4~mag brighter than a passively evolved $L_*$ galaxy at the same redshift. The luminosity of the host galaxy of the QSO would thus overlap with the highest found in radio-loud QSOs and radio-galaxies at the same redshift. ", "introduction": "Recent evidence that the cosmological evolution of the density of star formation in the Universe (Madau et al. 1996) follows closely the QSO density evolution (Boyle \\& Terlevich 1998) emphasizes the need to study the kinds of galaxies that host Active Galactic Nuclei in order to understand the link between star-formation and nuclear activity, and potentially the role of nuclear activity in galaxy formation. At the peak value of QSO density ($z\\approx 2$ to 3), the few QSO host-galaxies detected so far present rest-frame UV fluxes that reach up to 20\\%\\ of the total QSO luminosity, indicating star-formation rates about 200 \\Msun/yr and above for both radio-loud (Lehnert et al. 1992) and radio-quiet samples (Aretxaga, Boyle \\& Terlevich 1995, Hutchings 1995). These values are almost an order of magnitude above those of field galaxies at similar redshifts selected through Lyman Break techniques (Steidel et al. 1996, Lowenthal et al. 1997). The properties of these QSO hosts are not unprecedented, since they follow very closely the luminosity--size relation of nearby star forming galaxies, overlapping with its high-luminosity end (Aretxaga, Terlevich \\& Boyle 1998). However, the UV fluxes only carry information about the high-mass end of the stellar populations in the galaxies, and say little about the bulk of the stellar mass which is better characterized by optical to NIR observations. Although a few hosts of extreme radio-loud QSOs at $z\\approx 2$ have been detected in NIR bands (Lehnert et al. 1992, Carballo et al. 1998), attempts to image the hosts of normal radio-quiet QSOs at the same redshifts have been unsuccessful to date (Lowenthal et al. 1995, Aretxaga et al. 1998). Imaging radio-quiet systems, which constitute more than 95\\%\\ of all QSOs, is important in order to characterize the bulk of the population. The observed optical sizes of FWHM$\\approx 1$ arcsec (Aretxaga et al. 1995), clearly demand a technique which offers the highest available angular resolution. In this paper we focus our attention on the detection of the host of a normal radio-quiet $z\\approx 2$ QSO with the Adaptive Optics System in operation at the ESO 3.6m telescope in La Silla. Preliminary results on similar programs to image the host-galaxies of QSOs at $z\\approx 0.5 \\hbox{ \\ and \\ } 1.7$ using Adaptive Optics have been presented in a recent conference devoted to quasar hosts (Bremer et al. 1997, Hutchings 1997). \\ifoldfss ", "conclusions": "" }, "9804/astro-ph9804114_arXiv.txt": { "abstract": "We obtain self-similar solutions that describe the gravitational collapse of nonrotating, isothermal, magnetic molecular cloud cores. We use simplifying assumptions but explicitly include the induction equation, and the semianalytic solutions we derive are the first to account for the effects of ambipolar diffusion following the formation of a central point mass. Our results demonstrate that, after the protostar first forms, ambipolar diffusion causes the magnetic flux to decouple in a growing region around the center. The decoupled field lines remain approximately stationary and drive a hydromagnetic C-shock that moves outward at a fraction of the speed of sound (typically a few tenths of a kilometer per second), reaching a distance of a few thousand AU at the end of the main accretion phase for a solar-mass star. We also show that, in the absence of field diffusivity, a contracting core will not give rise to a shock if, as is likely to be the case, the inflow speed near the origin is nonzero at the time of point-mass formation. Although the evolution of realistic molecular cloud cores will not be exactly self similar, our results reproduce the main qualitative features found in detailed core-collapse simulations (Ciolek \\& K\\\"{o}nigl 1998). ", "introduction": "Low-mass stars are generally believed to form as a result of the gravitational collapse of molecular cloud cores. The cores are initially supported by thermal and magnetic forces, but because of ambipolar diffusion (the drift of ions, to which the magnetic field lines are attached, relative to the dominant neutral gas component), they gradually lose their magnetic support and eventually collapse after becoming ``supercritical'' (see, e.g., Mouschovias 1987 for a review).\\footnote{In this paper we reserve the term ``core'' for the high-density central region of a molecular cloud and do {\\em not} apply it to the point mass that forms from the collapse of such a core.} The most detailed numerical treatments to date of the problem of the ambipolar diffusion-initiated formation of supercritical cores and the early stages (prior to point mass formation) of their subsequent dynamical collapse have been presented by Mouschovias and collaborators (Fiedler \\& Mouschovias~1992, 1993; Ciolek \\& Mouschovias~1993, 1994, 1995, hereafter CM93, CM94, CM95; Basu \\& Mouschovias~1994, 1995a, 1995b, hereafter BM94, BM95a,b). Because the timescale for core formation is much longer than the timescale for dynamical collapse, special numerical techniques had to be employed in these calculations. The simulations were terminated when the central densities reached $\\sim 10^{10}\\ {\\rm cm}^{-3}$ and the underlying assumptions of isothermality (e.g., Gaustad~1963) and flux freezing onto the ions (e.g., Pneuman \\& Mitchell~1965) broke down. These calculations were nevertheless able to demonstrate that {\\em supercritical cores begin to collapse dynamically before a point mass (i.e., a protostar) appears at the origin}. The dynamical evolution of supercritical cores after their formation has been studied by many researchers. Solutions exist for the collapse of nonrotating, self-gravitating spheres without thermal support (Henriksen~1994), self-gravitating spheres with thermal support (Penston~1969; Larson~1969; Shu~1977; Hunter~1977; Boss \\& Black~1982; Whitworth \\& Summers~1985; Foster \\& Chevalier~1993) as well as with a combined thermal and isotropic magnetic pressure support (Chiueh \\& Chou 1994), and self-gravitating disks with thermal support (Narita, Hayashi, \\& Miyama~1984; Matsumoto, Hanawa, \\& Nakamura~1997) and also with ordered, frozen-in magnetic fields (Nakamura, Hanawa \\& Nakano~1995; Li \\& Shu~1997, hereafter LS). In order to choose a particular solution for a given problem, one needs to know the properties of the supercritical core at the time of its formation. This information, however, can only be gleaned from a study of the preceding, quasi-static evolution of the core under the influence of ambipolar diffusion. Although different assumptions about the initial state of the core yield solutions that are qualitatively similar in their gross behavior (the core collapses with near free-fall speeds and a point mass eventualy forms at the center), the solutions do differ in such important details as the accretion rate onto the central point mass and the formation (or absence) of shocks. The well-known examples of the Larson-Penston (1969) and Shu~(1977) similarity solutions in fact represent two extremes of a whole continuum of self-similar collapse solutions specified by a cloud's initial configuration and the conditions at its boundary (Hunter 1977; Whitworth \\& Summers 1985; see also Chiueh \\& Chou 1994 for a generalization to the case of an isotropic internal magnetic pressure). The Larson-Penston (1969) solution is characterized by a spatially uniform, supersonic (at $\\sim 3.3$ times the isothermal speed of sound $C$) infall speed and an inverse-square dependence of the density $\\rho$ on the radius $r$ at the instant of point-mass formation (PMF); the mass accretion rate at the center is $\\sim 29 \\ C^3/G$ (where $G$ is the gravitational constant) at that instant and increases to $\\sim 47 \\ C^3/G$ immediately after PMF. Numerical simulations of the collapse of nonmagnetic isothermal spheres (Hunter 1977; Foster \\& Chevalier 1993) have indicated that this solution provides a good approximation to the conditions near the center at the PMF epoch for clouds that are initially near a marginally stable equilibrium. The Shu (1977) solution strictly applies only to the post-PMF evolutionary phase: it consists of an inner free-fall region and a hydrostatic outer envelope that are separated by an outward-propagating (at a speed $C$ relative to the gas) expansion wave. The envelope corresponds to a singular isothermal sphere ($\\rho \\propto r^{-2}$) and the mass accretion rate onto the center is $\\sim 1 \\ C^3/G$. In applying this solution to real systems, it was proposed to identify the initial core configuration at the end of the quasi-static ambipolar-diffusion phase with a singular isothermal sphere (or, more generally, a toroid) at the instant of PMF (e.g., Shu, Adams, \\& Lizano 1987; Li \\& Shu 1996). However, as we noted above, the conclusion from detailed numerical simulations has been that the dynamical phase of core collapse generally commences well before the PMF epoch, so that the innermost region is not well represented by a quasi-static solution at the time of point-mass formation. Another interesting effect that depends on the specific choice of initial conditions and on the detailed physical properties of the collapsing core is the formation (or absence) of shocks (e.g., Tsai \\& Hsu 1995; LS). For example, LS discovered that when, instead of a spherical core, one considers the collapse of a flattened disk, the expansion wave of Shu~(1977) becomes a shock. As we show in this paper, when one takes proper account of the fact that supercritical cores collapse dynamically before a point mass first forms at the origin, that shock disappears. Nevertheless, a physical basis for the formation of shocks in collapsing magnetized molecular cloud cores has been discussed by Li \\& McKee (1996), who argued that a hydromagnetic C-shock will appear as a result of the outward diffusion of inwardly advected magnetic flux. The existence of such a shock has been confirmed in the numerical simulations of Ciolek \\& K\\\"{o}nigl (1998, hereafter CK), and it is, in fact, a salient feature of the semianalytic solutions derived in this paper. The aim of the present work is to clarify the effects of ambipolar diffusion in dynamically collapsing supercritical cores. Toward this goal, we construct semianalytic, time-dependent similarity solutions of gravitationally contracting, magnetized, isothermal disks. Although the evolution of real molecular cloud cores is not expected to be exactly self similar, we demonstrate, through a comparison with the detailed numerical simulations of CK, that our solutions capture the main traits exhibited by the latter calculations. Based on an analogous comparison with the results of numerical simulations, Basu (1997) showed that a self-similar scaling describes the pre-PMF evolution in the innermost flux tubes of collapsing supercritical cores quite well. To complement his study, we concentrate in this paper on the post-PMF evolutionary phase. Our approach differs, however, from that of Basu (1997) in that we explicitly solve the induction equation, whereas he accounted for the effects of ambipolar diffusion only in a phenomenological manner.\\footnote{Our work is thus also distinguished from that of Safier, McKee, \\& Stahler (1997), who studied the effects of ambipolar diffusion in the spherically symmetric, quasi-static limit without explicitly solving the induction equation.} In fact, the solutions that we derive, while involving various simplifications, are nevertheless the first to consistently incorporate ambipolar diffusion into a self-similar representation of the collapse of a magnetized cloud core. \\footnote{The effect of {\\em weak} magnetic fields on a dynamically collapsing core in the presence of ambipolar diffusion was previously investigated by Galli \\& Shu (1993a), who carried out a perturbation expansion of the (nonmagnetic) spherical similarity solution of Shu (1977). As was already noted and discussed by Li \\& McKee (1996), the semianalytic solution derived in that paper, as well as the associated numerical calculation in Galli \\& Shu (1993b), did not uncover the existence of a flux diffusion-driven shock.} We formulate the problem in \\S 2, present our solutions in \\S 3, and discuss the results in \\S 4. Our conclusions are summarized in \\S 5. ", "conclusions": "In this paper we have presented a self-similar solution of the collapse of a magnetized molecular cloud core (assumed to also be nonrotating and isothermal) that, for the first time, incorporated the effects of ambipolar diffusion in a self-consistent manner. We have focused on the post-PMF (point-mass formation) phase of the collapse of a disk-like core, noting that Basu (1997) had previously explored the self-similar nature of the collapse before a central mass (i.e., a protostar) first appears at the origin. We clarified the distinction between the ideal and nonideal MHD cases by plotting the singular lines in the position--velocity space and showing that they correspond to different critical speeds (the magnetosonic speed and thermal sound speed in the ideal and nonideal problems, respectively). We obtained a solution for the ideal (flux-frozen) case that exhibits a split-monopole field topology near the center. This solution differs from the one obtained by Li \\& Shu (1997) in that it involves no shocks. We showed that the shock in the LS solution is a direct consequence of their assumption that the core at the time of PMF is described by a stationary density distribution (corresponding to a singular isothermal toroid), and we pointed out that a shock will generally {\\em not} be present under the more realistic assumption of a nonzero inflow speed near the origin at that instant. We demonstrated, however, that a shock is a generic feature of the solution in the nonideal (ambipolar diffusion) case. This (C-type) shock is a direct consequence of the action of ambipolar diffusion in the central region of the core following PMF: the magnetic diffusivity decouples the field from the matter, causing the gas to free-fall to the center (where it accumulates in a point mass) and the field to stay behind and drive a shock outward. We have compared this solution with the results of the numerical simulations of Ciolek \\& K\\\"onigl (1998) and confirmed that, while the more realistic numerical models are not strictly self-similar, our simplified solution nevertheless captures the main features of the core evolution after PMF." }, "9804/astro-ph9804138_arXiv.txt": { "abstract": "s{ We review the topic of Cosmic Microwave Background Anisotropy measurements carried out by means of balloon-borne telescopes. After a short description of the experimental methodology, we outline the peculiar problems of these experiments, and we describe the main results obtained and the perspects for future developments. } ", "introduction": " ", "conclusions": "Ballooning for CMB Anisotropy measurements is a very active field worlwide. The activity is growing, for two main scientific reasons: 1) High frequency ($>$90 GHz) and high angular resolution ($\\sim$ 10 arcmin FWHM) measurements are possible and effective. 2) These measurements complement the forthcoming data from MAP and are a very important test-bed for Planck technologies. Moreover, good science is being produced, and promises for important results (like the $\\ell$-space spectroscopy of the acoustic peaks, or detection/falsification of non-gaussian statistics for the CMB fluctuations) are quite convincing. As an example relevant for this conference, we can mention the fact that determination of several cosmological parameters is possible with very good accuracy from LDB experiments. For example, if all the systematics effects are properly removed, a single LDB experiment with 12 arcmin FWHM beams, 16 total power bolometric detectors with sensitivity of 80 $\\mu K \\sqrt{s}$, 10 days of observing time spent over a $50^o \\times 50^o$ sky region, can measure the power spectrum of the CMB anisotropies with very good accouracy \\cite{Silvia}. Fits can be done on the measured power spectrum \\cite{KL}, allowing to recover $\\Omega_{tot}$ with a 3$\\%$ error, $\\Omega_{\\Lambda}$ with 6$\\%$ error, $\\Omega_B$ with $1\\%$ error, $n$ scalar with 18$\\%$ error, $H_o$ with 1$\\%$ error, $Q_{rms,PS}$ with 4$\\%$ error. Here the errors for any parameter make no assumptions about the value of the other parameters. These measurement errors can be significantly reduced if one or more of the cosmological parameters are constrained by other observations or fixed by assumptions." }, "9804/astro-ph9804281_arXiv.txt": { "abstract": "We present a description of the observations and data reduction procedures for an extensive spectroscopic and multi-band photometric study of nine high redshift, optically-selected cluster candidates. The primary goal of the survey is to establish new constraints on cluster and galaxy evolution, with specific emphasis on the evolution of galaxy morphology and on the star-formation history of the galaxies within and around distant clusters. We have measured 892 new redshifts for galaxies with $R \\le 23.3$. The data will also serve as deep probes of the foreground and background large-scale structures. The observations include broad band optical imaging and spectroscopy with the Low Resolution Imaging Spectrograph at the 10 meter W. M. Keck Observatory telescope; K-band imaging with IRIM at the 4 meter Kitt Peak National Observatory telescope; and deep, high angular resolution imaging with the WFPC2 onboard the Hubble Space Telescope. We also describe the procedures used to obtain morphological information. We have established that six of the nine cluster candidates are indeed real space density enhancements and are representative of those typically associated with clusters of galaxies. The remaining three candidates appear to be projections of several smaller groups at widely separated distances. This success rate is consistent with estimates of the false positive rate in 2D optical high-$z$ cluster searches. ", "introduction": "Clusters of galaxies have historically provided an important tool for studying cosmology and the evolution of galaxies. Because of their high concentration of galaxies, clusters provide an environment in which to study large, statistical samples of galaxies. Therefore, examining clusters of galaxies from the local universe to those at high redshift allows us to probe galactic evolution to redshifts of the order of 1. Clusters of galaxies at redshifts of $z \\simless 0.2$ have been extremely well cataloged in the optical regime (e.g.\\ Abell 1958; Zwicky \\etal 1968; Dressler 1980; Shectman 1985; Abell \\etal 1989; Lumsden \\etal 1992; Dalton \\etal 1994). The analyses of local clusters indicate that these systems are dense (Abell richnesses of $\\sim 30 - 300$ galaxies), massive ($M \\sim 10^{14} - 2 \\times 10^{15}~h^{-1}~{\\rm M_{\\odot}}$), and dominated by early-type galaxies ($\\sim 50 - 80\\%$ of the total galaxy population). These studies provide a strong basis on which to compare cluster properties at increasingly higher redshift. Detailed photometric, spectroscopic and morphological studies have been extended to clusters of galaxies at redshifts up to $z \\sim 0.6$. The first substantial contribution at these redshifts came from Butcher \\& Oemler (1984) who found a surprisingly large population of blue galaxies in conjunction with the expected red sequence of early-type cluster members. Further photometric and spectroscopic campaigns, including the ambitious CNOC and MORPHS surveys, have confirmed the progressive bluing of the cluster's galaxy population and have tracked the passive evolution of the early-type galaxies (Dressler \\& Gunn 1992; Oke, Gunn \\& Hoessel 1996; Yee, Ellingson \\& Carlberg 1996; Ellingson \\etal 1997; Ellis \\etal 1997; Stanford et al.\\ 1995, 1997; and references therein). The Hubble Space Telescope (HST) has enabled morphological classification of intermediate-redshift ($z \\simless 1$) galaxies on scales which are comparable to the classifications made of their local counterparts. These high-resolution studies have revealed that there may be a substantial change in the morphological composition of the clusters (Smail \\etal 1997; Dressler \\etal 1997; however, see Stanford \\etal 1997). All of these results imply that there is a significant amount of evolution occurring in the cluster environment between redshifts of $z \\approx 0.5$ and $z = 0.0$. In order to understand this apparent change in the galaxy population, it is essential to probe in similar detail clusters of galaxies at even higher redshift where the effects of evolution and cosmology will be even greater. In light of this, we have undertaken an extensive survey of nine candidate clusters of galaxies at redshifts greater than 0.6. Only a few optical/near-IR surveys have attempted to detect systematically clusters at high redshift; therefore, we have chosen our cluster sample from the Gunn, Hoessel, \\& Oke (1984) survey and the Palomar Distant Cluster Survey (Postman \\etal\\ 1996). The observational data compiled for this survey, to date, includes deep $BVRIK$ photometry, over 900 low-resolution spectra, and deep F606W/F702W/F814W imagery from HST. The large redshift database allows us to reliably distinguish between physically real clusters and chance line-of-sight projections. The full data allow us to measure the global properties of the clusters, such as profile shape and dynamics, as well as the individual properties of the cluster galaxies, such as color, star-formation rate, and morphology. In this introductory paper to our high-redshift cluster series, we describe in detail the observational and data reduction techniques of each aspect of this survey and present the redshift histograms for our nine fields. The subsequent papers in this series will present the specific analyses and scientific results of this survey. These papers include the second and third installments in this series which present a detailed photometric, spectroscopic, and morphological analyses of the first two clusters to be completed in this survey, CL0023+0423 and CL1604+4304 (Postman, Lubin \\& Oke 1998; Lubin \\etal 1998). ", "conclusions": "We have described the data acquisition and reduction procedures of our photometric and spectroscopic campaign to study nine candidate clusters of galaxies at redshifts of $z > 0.6$. The observational program consists of four main parts : \\newcounter{discnt} \\begin{list} {\\arabic{discnt}.} {\\usecounter{discnt} \\setlength{\\leftmargin 0.2in}{\\itemsep 0in}{\\topsep 0in}{\\parskip 0in}} \\item Spectra for approximately 80\\% of all galaxies down to a Johnson--Cousins $R$ magnitude of $\\sim 23.5$ within a fixed area of ${2}^{'}.2 \\times {7}^{'}.6$ of each cluster field using the Low--Resolution Imaging Spectrograph (LRIS) at the Keck 10m telescope. We have obtained spectra covering the range 4400 \\AA\\ to 9500 \\AA\\ for $\\sim 130 - 150$ galaxies per cluster field. Redshifts have been determined for 892 objects. \\item Deep $BVRI$ imaging with Keck of all galaxies in the full LRIS field of $6^{'} \\times 8^{'}$ centered on each cluster. The $5\\sigma$ detection limits are $B = 25.1$, $V = 24.1$, $R = 23.5$, and $I = 21.7$ in our standard 3 arcsecond radius aperture. The photometric data are converted to absolute fluxes in order to obtain absolute spectral energy distributions. \\item High angular resolution imagery with HST in order to provide morphological information on the galaxies in the WFPC2 field-of-view (${160}^{''} \\times {160}^{''}$) centered on each cluster. Each of the cluster candidates has been or will be observed by HST in Cycles 5 and 6. \\item High precision $K$ band photometry across the WFPC2 field--of--view with the KPNO 4m telescope for each cluster. The infrared survey reaches a limiting magnitude of $K^{'} = 20$ in the standard aperture. \\end{list} We have presented the redshift histograms for the nine candidate clusters of galaxies in this survey. We find that six of the nine candidate clusters of galaxies are real density enhancements. They include CL0023+0423, CL0943+4804, CL1324+3011, CL1325+3009, CL1604+4304, and CL1604+4321. The remaining three candidates are of a more dubious nature. Their redshift distributions reveal no clear density enhancement but rather an apparent superposition of small groups of galaxies along the line-of-sight. This false positive rate is consistent with the estimate of $\\sim 30$\\% provided in Postman \\etal (1996). At lower redshifts ($z \\simless 0.5$) the spurious rate is about 20\\% or less. This is based on spectroscopic follow-up of PDCS candidates being done by Holden \\& Nichol (1998) at the KPNO 4m telescope. We conclude that optical detection of clusters remains a successful and important method for identifying such systems out to $z \\sim 1$ and, further, will provide an important complement to cluster searches at other wavelengths. Results on the star-formation history, dynamics, and morphological properties of CL0023+0423 ($z = 0.84$) and CL1604+4304 ($z = 0.90$) are presented in Papers II and III. \\vskip 1.0cm We thank Don Schneider and the anonomous referee for their invaluable comments on this manuscript. The W.M. Keck Observatory is operated as a scientific partnership between the California Institute of Technology, the University of California, and the National Aeronautics and Space Administration. It was made possible by the generous financial support of the W. M. Keck Foundation. LML graciously acknowledges support from a Carnegie Fellowship. Support for this work was also provided, in part, by NASA through grant number GO-06000.01-94A from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS5-26555. \\clearpage" }, "9804/astro-ph9804004_arXiv.txt": { "abstract": "s{We review estimates of small scale fluctuations due to extragalactic point sources in the Planck Surveyor frequency bands. While our undestanding of the spectral and evolutionary properties of these sources is far from complete, conservative estimates allow us to confidently conclude that, in the frequency range 100--200 GHz, their contaminating effect is well below the expected anisotropy level of the cosmic microwave background (CMB), down to angular scales of at least $\\simeq 10'$. Hence, an accurate subtraction of foreground fluctuations is not critical for the determination of the CMB power spectrum up to multipoles $\\ell \\simeq 1000$. In any case, Planck's wide frequency coverage will allow to carefully control foreground contributions. On the other hand, the all sky surveys at 9 frequencies, spanning the range 30--900 GHz, will be unique in providing complete samples comprising from several hundreds to many thousands of extragalactic sources, selected in an essentially unexplored frequency region. New classes of sources may be revealed in these data. The familiar ``flat''-spectrum radio sources should show spectral features carrying essential information on their physical properties. Crucial information will be provided to understand the nature of radio sources with strongly inverted spectra. Scenarios for the cosmological evolution of galaxies will be extensively tested.} ", "introduction": "The multifrequency all-sky maps produced by the Planck Surveyor mission will comprise, in addition to anisotropies which are outgrowths of primordial fluctuations, and whose precision measurements are the main goal of the mission, astrophysical foregrounds, the most important of which, over the frequency range of interest, are those due to emissions in our own Galaxy and to extragalactic radio and mm/sub-mm sources. We deal here with extragalactic sources, which may be a major limiting factor for experiments, like Planck, aimed at accurately determining the cosmic microwave Background (CMB) power spectrum $C_\\ell$ up to multipoles $\\ell \\sim 2000$, corresponding to angular scales $\\theta \\sim 5'$. In fact, a Poisson distribution of sources produce a white noise power spectrum with the same power in all multipoles~\\cite{Tegmark}, so that their contribution to fluctuations in a unit logarithmic multipole interval increases with $\\ell$ as $\\ell(\\ell +1)C_\\ell \\propto \\ell^2$ (for large values of $\\ell$), while, at least for the standard inflationary models, which are consistent with the available anisotropy detections, the function $\\ell(\\ell +1)C_\\ell$ yielded by primordial CMB fluctuations is approximately constant for $\\ell \\lsim 100$, then oscillates and finally decreases quasi exponentially for $\\ell \\gsim 1000$ ($\\theta \\lsim 10'$). Hence confusion noise due to discrete sources will dominate at small enough angular scales. In \\S$\\,$2 we summarize the limitations set by fluctuations due to extragalactic sources on Planck measurements of primordial CMB anisotropies. On the other hand, the multifrequency all sky surveys carried out by the Planck Surveyor mission will provide a very rich database for astrophysical studies; their impact on investigations of physical and evolutionary properties of different classes of extragalactic sources is briefly outlined in \\S$\\,$3. Our main conclusions are presented in \\S$\\,$4. ", "conclusions": "Luckily enough, both for galaxies and active galactic nuclei, the crossover between the radio and the dust emission components, determining a minimum in the spectral energy distribution, is roughly coincident with the CMB intensity peak. The dust temperature tends to be higher for bright distant objects, moving the minimum to higher frequencies in the rest frame and thus partially compensating for the effect of redshift. This situation makes the mm region ideal for mapping primordial anisotropies. Although our understanding of foregrounds at Planck frequencies is far from complete, estimates using worst-case parameters in extrapolating existing measurements to Planck frequencies or angular scales, allow us to safely conclude that, in the frequency range 100-200 GHz, the foreground fluctuations, which are dominated, on small scales ($\\theta \\lsim 30'$), by extragalactic sources, are well below the expected amplitude of CMB anisotropies over much of the sky. Hence, the removal of foreground contamination is not critical for accurate determinations of the power spectrum of CMB anisotropies up to multipoles of at least $\\ell \\sim 1000$. On the other hand, while only a small fraction of high Galactic latitude pixels are strongly contaminated by astrophysical foregrounds, the Planck surveys at 9 frequencies will provide sufficiently rich complete samples for astrophysical studies. Spectral information will be provided for ``flat''-spectrum radio sources (compact radio galaxies, radio loud QSOs, BL Lacs, blazars) over a frequency region where spectral features carrying essential information on their physical conditions show up (breaks due to energy losses of relativistic electrons, self-absorption turnovers of flaring components, ...). Planck surveys will be unique in providing complete samples of bright radio sources with inverted spectra, essentially undetectable in radio-frequency surveys. Important classes of sources of this kind are GHz peaked spectrum sources, which may be the youngest stages of radio source evolution and may thus provide insight into the genesis of radio sources, and advection dominated sources, corresponding to final stages of accretion in giant elliptical galaxies hosting a massive black hole. The high frequency Planck channels will detect thousands of dusty galaxies, a large fraction of which at substantial redshifts, allowing to extensively test scenarios for galaxy evolution. The increasing evidence that a large, and perhaps dominant fraction, of star formation at high redshifts may be hidden by dust, makes far-IR to sub-mm surveys an essential complement to optical data." }, "9804/astro-ph9804142_arXiv.txt": { "abstract": "Using the IRAM 30-m telescope, we observed the supernova remnant 3C~391 (G31.9+0.0) and its surroundings in the \\cotwo, \\hcop, \\cstwo, \\csthree, and \\csfive\\ lines. The ambient molecular gas at the distance (9 kpc) of the remnant comprises a giant molecular cloud whose edge is closely parallel to a ridge of bright non-thermal radio continuum, which evidently delineates the blast-wave into the cloud. We found that in a small (0.6 pc) portion of the radio shell, the molecular line profiles consist of a narrow (2 \\kms) component, plus a very wide ($> 20$ \\kms) component. Both spectral components peak within $20^{\\prime\\prime}$ of a previously-detected OH 1720 MHz maser. We name this source 3C~391:BML (broad molecular line); it provides a new laboratory, similar to IC 443 but on a larger scale, to study shock interactions with dense molecular gas. The wide spectral component is relatively brighter in the higher-excitation lines. We interpret the wide spectral component as post-shock gas, either smoothly accelerated or partially dissociated and reformed behind the shock. The narrow component is either the pre-shock gas or cold gas reformed behind a fully dissociative shock. Using the 3 observed CS lines, we measured the temperature, CS column density, and H$_2$ volume density in a dense clump in the parent molecular cloud as well as the wide-line and narrow-line portions of the shocked clump. The physical conditions of the narrow-line gas are comparable to the highest-density clumps in the giant molecular cloud, while the wide-line gas is {\\it both} warmer and denser. The mass of compressed gas in 3C~391:BML is high enough that its self-gravity is significant, and eventually it could form one or several stars. ", "introduction": "Supernovae are thought to be the source of kinetic energy of the interstellar medium, keeping the gas in motion and returning material from dense molecular clouds into the more diffuse interstellar medium and the galactic halo. When a massive star ends its life in a supernova explosion, it often does so in the vicinity of the molecular cloud in which it was born, as is evidenced by the close correspondence of OB associations and giant H~II regions in spiral arms (\\cite{elmlad77}). Despite the expected close association between Type II supernovae and molecular clouds, very few cases of supernova-molecular cloud (SN-MC) interaction are known or suspected. The blast wave from a supernova within or near the edge of a cloud will progress rapidly through the inter-clump medium and drive slower shocks into dense clumps. Multiple reflections of high-energy charged particles within the complicated magnetic field of an SN-MC interaction are a possible source of cosmic rays, which will permeate the entire region (\\cite{chevalier77}; \\cite{esposito96}). The thermal radiation from the remnant interior (mainly X-rays), cosmic rays and their secondary gamma rays, and direct impact of the blast wave onto clumps should visibly perturb the excitation, chemistry, and dynamics of the parent molecular cloud for at least the $\\sim 10^5$ year period during which the SN blast wave is most powerful. So far, the only well-known case of an SN-MC interaction is IC 443, where molecular lines have been detected with FWHM $\\sim$ 20 km s$^{-1}$ (much wider than the lines from nearby, un-shocked gas), and from energy levels far above the ground state (\\cite{white87}; van Dishoeck, Jansen, \\& Phillips 1993; \\cite{wang92}). Other remnants, including W~28, CTB 109, Kes 79, and W~51C have been suggested as SN-MC interactions based on their proximity to molecular clouds, wide molecular lines, or both (\\cite{woot77}; \\cite{woot81}; \\cite{tatematsu90}; \\cite{green92}; \\cite{koo97a}; \\cite{koo97b}). In the case of W~28, W~44 and 3C~391, 1720 MHz OH emission has been detected from many small spots, with brightness temperatures so high that they must be masers; these masers are thought to be collisionally excited and they strongly suggest the presence of SN-MC interactions (Frail, Goss, \\& Slysh 1994; \\cite{frail96}). 3C~391 is one of the brightest radio supernova remnants, and high-resolution radio images suggest a `break-out' morphology due to an explosion near the edge of a molecular cloud (\\cite{rm93}). The X-ray emission from 3C~391 peaks in its interior and has a thermal spectrum (\\cite{rp96}), characteristic of a newly-defined class of supernova remnants, called `mixed-morphology' remnants, whose nature has been linked to interaction with a strongly inhomogeneous pre-shock interstellar medium (\\cite{rp98}). A recent map of the \\coone\\ emission in the vicinity of 3C~391 revealed a giant molecular cloud that is precisely parallel to the bright ridge of radio emission, confirming that its `break-out' radio morphology is indeed due to the strong density contrast between the molecular cloud to the northwest and the relatively empty regions elsewhere (Wilner, Reynolds, \\& Moffett 1998). The work described in this paper is part of our recently-initiated campaign to search for and characterize SN-MC interactions in the mixed-morphology supernova remnants. Our first result was the detection of bright [O~I] 63 $\\mu$m and dust emission from 3C~391, showing that the blast-wave into the molecular gas is radiative, and the SN-MC interaction is a significant energy loss for the remnant, although it remains globally adiabatic (\\cite{reach96}). In this paper, we present new observations of molecular emissions from 3C~391, designed to search for the effects of the SN-MC interaction on the molecular cloud, using millimeter-wave observations at high angular resolution and several transitions requiring a range of physical conditions for excitation. Throughout this paper, we assume a distance to 3C~391 of 9~kpc, which is based on the comparison of H~I 21-cm emission and absorption line profiles (\\cite{radakrish}) and the H$_2$CO absorption line at 96 \\kms\\ (\\cite{downes}); our adopted distance is consistent with that adopted by others for this remnant (\\cite{rm93}). ", "conclusions": "We observed the entire supernova remnant 3C~391 in millimeter lines of CS, CO, and HCO$^+$. The lower-excitation lines reveal a giant molecular cloud to the northwest of the remnant, explaining the `break-out' morphology of the radio emission. The interactions between the blast wave and a very dense molecular clump was found within a small ($50^{\\prime\\prime}$) region that we call 3C~391:BML. A wide component (FWHM 25 \\kms) and a narrow component (FWHM 2 \\kms) both peak at 3C~391:BML, which is coincident with an OH 1720 MHz maser. The excitation of the wide molecular lines requires both high gas temperature ($> 100$ K) and density ($\\sim 3\\times 10^5$ cm$^{-3}$). The narrow-line region require somewhat lower density and are consistent with much lower ($\\sim 20$ K) temperatures. We identified a clump in the parent molecular cloud with properties similar to the narrow-line region. Therefore, the 3C~391:BML clump was similar to the highest-density clumps in the parent molecular cloud, and it is currently being shocked. The brightness of the wide \\cotwo\\ line from 3C~391:BML is consistent with C-type molecular shocks with $10^4 < n_0 < 10^5$ cm$^{-3}$ and $10 < v_s < 50$ \\kms, or J-type shocks with $n_0\\sim 10^3$ and $v_s \\sim 100$ \\kms. The pressure in the shocked clump is much higher than the estimated ram pressure of the remnant, possibly because of its self-gravity. If so, this clump is a likely site of triggered star formation. A widespread interaction of 3C~391 with molecular gas is evidenced by the distribution of CO and CS lines with central velocities offset from that of the parent cloud by 10 to 15 \\kms; these velocities correspond to those of the two OH masers. The interaction comprises nearly an entire hemisphere of the remnant, making 3C~391 a `CO shell' remnant. {\\bf Acknowledgment} We thank Hans Ungerechts at IRAM for helping us get started on the IRAM telescope and David Wilner and Steve Reynolds for sharing early results of their observations. WTR thanks the Commissariat d'Energie Atomique, in Saclay, France for hospitality and computing power during part of the data analysis. We thank Bon-Chul Koo for his comments and support. The research described in this paper was carried out in part by the California Institute of Technology, under a contract with the National Aeronautics and Space Administration. \\clearpage \\begin{deluxetable}{lllllll} \\footnotesize \\tablecaption{Observed spectral lines and Telescope parameters\\label{tab:telparams}} \\tablewidth{0pt} \\tablehead{ \\colhead{transition} & \\colhead{frequency} & \\colhead{$T_{sys}$\\tablenotemark{a}} & \\colhead{$\\eta_{mb}$} & \\colhead{beam} & \\colhead{$\\delta v$\\tablenotemark{b}} & \\colhead{$\\Delta v$\\tablenotemark{c}} \\\\ & \\colhead{(GHz)} & \\colhead{(K)} & & \\colhead{($^{\\prime\\prime}$)} & \\colhead{(\\kms)} & \\colhead{(\\kms)} } \\startdata HCO$^+$($1\\rightarrow 0$) & 89.1885 & 310 & 0.82 & 27 & 0.22 & 430\\nl CS($2\\rightarrow 1$) & 98.9798 & 250 & 0.76 & 24 & 0.24 & 430\\nl CS($3\\rightarrow 2$) & 146.9690 & 240 & 0.58 & 16 & 0.65 & 290\\nl CS($5\\rightarrow 4$) & 244.9356 & 490 & 0.43 & 10 & 1.2 & 310\\nl CO($2\\rightarrow 1$) & 230.5380 & 720 & 0.45 & 10 & 1.3 & 330\\nl \\enddata \\tablenotetext{a}{typical system temperature for observations presented in this paper} \\tablenotetext{b}{velocity resolution} \\tablenotetext{c}{velocity coverage} \\end{deluxetable} \\clearpage \\begin{deluxetable}{llll} \\tablecolumns{4} \\footnotesize \\tablecaption{Measured properties of spectral lines\\tablenotemark{a}\\label{tab:spectab}} \\tablewidth{0pt} \\tablehead{ \\colhead{transition} & \\colhead{$T_{mb}$} & \\colhead{$V_{LSR}$} & \\colhead{$\\Delta V$ (FWHM)} \\\\ & \\colhead{(K)} & \\colhead{(\\kms)} & \\colhead{(\\kms)} } \\startdata \\cutinhead{wide-line position in shocked clump $(-40^{\\prime\\prime},-50^{\\prime\\prime})$\\tablenotemark{b}} HCO$^+$($1\\rightarrow 0$) & 0.91 & 111.4 & 25.5 \\nl CS($2\\rightarrow 1$) & 0.32 & 108.9 & 19.0 \\nl & 0.24 & 104.2 & 1.4 \\nl CS($3\\rightarrow 2$)\t & 0.37 & 108.9 & 20.0 \\nl CS($5\\rightarrow 4$)\t & 0.35 & 108.5 & 15.6 \\nl CO($2\\rightarrow 1$)\t & 17.6 & 111.1 & 22.5 \\nl & 15.1 & 104.2 & 2.6 \\nl \\cutinhead{narrow-line position in shocked clump $(-10^{\\prime\\prime},-85^{\\prime\\prime})$\\tablenotemark{b} } HCO$^+$($1\\rightarrow 0$) & 1.0 & 105.3 & 1.7 \\nl & 0.38 & 102.6 & 14.4 \\nl CS($2\\rightarrow 1$) & 0.90 & 105.5 & 1.0 \\nl & 0.23: & 103.7 & 6.0 \\nl CS($3\\rightarrow 2$)\t & 0.55 & 105.4 & 1.2 \\nl & 0.23 & 103.7 & 6.0 \\nl CS($5\\rightarrow 4$)\t & $<0.13$ & & \\nl CO($2\\rightarrow 1$)\t & 11.6 & 105.5 & 1.2 \\nl & 9.7 & 104.8 & 3.7 \\nl \\cutinhead{radio ridge $(-130^{\\prime\\prime},+90^{\\prime\\prime})$\\tablenotemark{c} } HCO$^+$($1\\rightarrow 0$) & 0.22: & 96.4 & 2.4 \\nl CS($2\\rightarrow 1$) & $<0.03$ & & \\nl CS($3\\rightarrow 2$) & $<0.04$ & & \\nl CS($5\\rightarrow 4$) & $<0.13$ & & \\nl CO($2\\rightarrow 1$)\t & 8.9 & 96.5 & 4.0 \\nl & 3.3 & 102.0 & 2.1 \\nl & 3.6 & 107.0 & 3.8 \\nl \\cutinhead{clump in parent cloud $(-130^{\\prime\\prime},+240^{\\prime\\prime})$\\tablenotemark{c} } HCO$^+$($1\\rightarrow 0$) & 0.32 & 96.1 & 7.3: \\nl CS($2\\rightarrow 1$) & 0.55 & 96.7 & 3.8 \\nl CS($3\\rightarrow 2$)\t & 0.27 & 96.5 & 3.6 \\nl CS($5\\rightarrow 4$)\t & $<0.13$ & & \\nl CO($2\\rightarrow 1$)\t & 12.0 & 96.0 & 5.4 \\nl & 8.4 & 107.9 & 4.0 \\nl \\enddata \\tablenotetext{a}{only components between 90 and 140 \\kms\\ are listed} \\tablenotetext{b}{spectra averaged within $11^{\\prime\\prime}$ radius} \\tablenotetext{c}{spectra averaged within $21^{\\prime\\prime}$ radius} \\end{deluxetable} \\clearpage" }, "9804/astro-ph9804156_arXiv.txt": { "abstract": "The scattering diameters of \\sgra\\ and several nearby OH masers ($\\approx 1\\arcsec$ at 1~GHz) indicate that a region of enhanced scattering is along the line of sight to the Galactic center. The scattering diameter of an extragalactic source seen through this scattering region will be larger by the ratio of the Sun-GC distance to the GC-scattering region separation. This ratio could be a factor of a few, if the scattering region is far from the GC and only a random superposition with it, to more than 100, if the scattering region is within the \\hbox{GC}. We have used the VLA to survey 10 (11) fields at 20~cm (6~cm) that are between 7\\arcmin\\ and 137\\arcmin\\ from \\sgra. Our objective was to identify extragalactic sources and measure their scattering diameters so as to constrain the GC-scattering region separation. In order to find sources within these fields, we have employed pdf\\clean, a source detection algorithm in which sources are identified in an image by comparing the intensity histogram of the image to that expected from a noise-only image. We found over 100 sources, with the faintest sources being approximately 3~mJy. The average number of sources per field is approximately 10, though fields close to \\sgra\\ tend to contain fewer sources. In a companion paper we combine our survey with previous observations of the GC, and we assess the likelihood that the scattering region is so close to the GC that the resulting scattering diameters cause extragalactic sources to be resolved out by our observations. A number of Galactic sources is included in our source catalog. We discuss the double-lobed source 1LC~359.872$+$0.178, potentially an X-ray quiet version of 1E~1740.7$-$2942, a shell-like structure with a central point source, and a possible radio transient. ", "introduction": "\\label{sec:gc.intro} If viewed through a plasma containing density fluctuations, an otherwise unresolved source will have a visibility, as measured by an interferometer of baseline length~$b$, of \\begin{equation} V(b) = \\exp\\left[-\\frac{1}{2}D_\\phi(b)\\right]. \\label{eqn:visibility} \\end{equation} The phase structure function, $D_\\phi(b) \\equiv \\langle[\\phi(0)-\\phi(b)]^2\\rangle$, is a measure of the phase perturbations, on a length scale~$b$, imposed on a propagating electromagnetic wave by fluctuations in the electron density. For a plane wave impinging on this scattering region \\begin{equation} D_\\phi(b) = 8\\pi^2r_{\\mathrm{e}}^2\\lambda^2\\int_0^D dz\\,\\int_0^\\infty dq\\,q[1 - J_0(bq)]P_{\\delne}(q, z), \\label{eqn:structurefunction1} \\end{equation} where $r_{\\mathrm{e}}$ is the classical electron radius, $J_0(x)$ is the zeroth-order Bessel function, $P_{\\delne}$ is the spatial spectrum of the density fluctuations, and the integral over $z$ is taken \\emph{from the source to the observer}. If the source of radiation is close to or embedded within the scattering medium, so that the medium is illuminated by spherical wavefronts, the argument of the Bessel function is $bq(z/D)$ (\\cite{i78}); the factor~$z/D$ accounts for the divergence of spherical waves. The apparent angular diameter of the source is determined by the width of the visibility function, and, hence, by how quickly $D_\\phi(b)$ decreases as a function of~$b$. Since $z/D < 1$, the difference in the form of the phase structure function for plane and spherical wavefronts means that sources close to the medium will show smaller angular diameters than those far from it. Hence, by comparing the scattering diameters of Galactic and extragalactic sources along similar lines of sight, one can constrain the \\emph{radial} location of the scattering material. Toward the Galactic center, the observed diameter of \\sgra\\ scales as $\\lambda^2$ over the wavelength range 30~cm to~3~mm (Davies, Walsh, \\& Booth~1976; \\cite{rogersetal94}), as expected if very strong interstellar scattering from microstructure in the electron density determines the observed diameter. Maser spots in OH/IR stars within 25\\arcmin\\ of \\sgra\\ also show enhanced angular broadening (\\cite{vfcd92}; \\cite{fdcv94}). The scattering disks of \\sgra\\ and many of the OH masers are observed to be anisotropic as well (\\cite{vfcd92}; \\cite{bzkrml93}; \\cite{krichbaumetal93}; \\cite{fdcv94}; \\cite{y-zcwmr94}); in the case of \\sgra, its scattering disk is anisotropic at least over the wavelength range 21~cm to 7~mm. These observations indicate that a region of enhanced scattering with an angular extent of at least 25\\arcmin\\ (60~pc at 8.5~kpc) is along the line of sight to \\sgra. At 1~GHz the level of angular broadening produced by this scattering region is roughly 10 times greater than that predicted by a recent model for the distribution of free electrons in the Galaxy (Taylor \\& Cordes~1993, hereinafter \\cite{tc93}), even though this model includes a general enhancement of scattering toward the inner Galaxy. Because all of the sources observed through this region have thus far been Galactic sources, with (presumably) approximately the same location (i.e., in the Galactic center), the radial location of the scattering region is unconstrained. The scattering region could be local to the Galactic center, within approximately 100~pc from the Galactic center---which we refer to as the GC model---or the region could be a random superposition and more than 1~kpc from the GC---which we refer to as the RS model. In the GC model, the region would be a site of excess scattering, and presumably arises from processes unique to the GC; in the RS model, the level of scattering in the region would be high, but not unusually so. Previous estimates for the location of the scattering region have ranged from 10~pc to 3~kpc. Ozernoi \\& Shisov~(1977) concluded that an ``unrealistic'' level of turbulence is implied unless the region is within 10~pc of the \\hbox{GC}. The level of turbulence they considered unrealistic, however, namely $\\sqrt{\\langle n_{\\mathrm{e}}^2\\rangle}/\\langle n_{\\mathrm{e}}\\rangle \\sim 1$, does appear to occur elsewhere in the interstellar medium (\\cite{s91}). Further, van~Langevelde et al.~(1992) showed that the free-free absorption toward \\sgra\\ would be excessive unless the scattering region was at least 0.85~kpc from the GC, though suitable adjustment of free parameters (outer scale and electron temperature) can decrease the limit to 0.03~kpc. With the free-free absorption they also placed an upper limit on the region's distance from the GC of 3~kpc. Although the GC model is attractive for phenomenological reasons, other sites of enhanced interstellar scattering are found throughout the Galaxy (e.g., NGC~6634, \\cite{mrgb90}; Cyg~X-3, \\cite{mmrj95}) and the mean free path for encountering such a region is approximately 8~kpc (\\cite{cwfsr91}). Identifying the location of the scattering region may provide clues to the origin of the scattering. The density fluctuations responsible for interstellar scattering are believed to be generated by velocity or magnetic field fluctuations (\\cite{h84}, 1986; Montgomery, Brown, \\& Matthaeus~1987; \\cite{s91}; \\cite{sg94}; \\cite{gs95}). Velocity or magnetic field fluctuations are also a natural means for inducing anisotropy in the density fluctuations and thereby in the scattering disks. If this supposition is correct, the amplitude of the density fluctuations may provide a measure of the coupling between the density and velocity or magnetic field fluctuations or, more generally, provide information about the small-scale velocity or magnetic field in the scattering region. However, because the radial location of the scattering region is unconstrained, relevant quantities, e.g., the rms density, are uncertain by a factor of $\\delgc/\\dgc$, where $\\delgc$ is the GC-scattering region separation and $\\dgc$ is the GC-Sun distance. Observations of extragalactic sources viewed through the scattering region could constrain $\\delgc$; however, few extragalactic sources have been identified toward the \\hbox{GC}. The two sources closest to \\sgra\\ are B1739$-$298 (\\cite{dkvgh83}) and GPSR~0.539$+$0.263 (Bartel~1994, private communication), which are 48\\arcmin\\ and 40\\arcmin\\ from \\sgra, respectively. Neither of these is within the region of enhanced scattering defined by the OH masers. This paper reports VLA and VLBA observations of potential extragalactic sources seen through the \\hbox{GC}. Section~\\ref{sec:gc.observe} describes the observations and data reduction, Section~\\ref{sec:catalog} discusses the identification of potential extragalactic sources and presents the catalog of sources, Section~\\ref{sec:sources} discusses certain Galactic sources found in our VLA survey, and Section~\\ref{sec:gc.conclude} discusses our results and presents our conclusions. A companion paper (Lazio \\& Cordes~1998, hereinafter \\cite{lc98}) combines the results of this paper with the previous observations of OH and \\hoh\\ masers and free-free emission in a likelihood analysis that constrains the angular extent and radial location of the scattering region. \\cite{lc98} also discusses the physical conditions inside the scattering region. ", "conclusions": "\\label{sec:gc.conclude} This paper has reported the results of a program to identify and obtain scattering diameters for extragalactic sources seen through the Galactic center scattering region. Because they are located far behind the GC, the scattering diameters of extragalactic sources, when compared to the scattering diameter of GC sources such as \\sgra, can constrain the \\emph{radial} location of the scattering region, viz.\\ equation~(\\ref{eqn:xgalsize}) and Figure~\\ref{fig:xgalsize}. Using the VLA we observed 10 (11) fields at~20~cm (6~cm) containing 15 suspected extragalactic sources. We increased our catalog of sources to well over 100 through the use of pdf\\clean: The intensity histogram of the primary beam was used to identify positive brightness image pixels that produced deviations from the shape of the expected noise-only histogram. We found approximately 10 sources per field. Follow-up VLBI observations on a subset of these sources have determined the scattering diameters for two heavily scattered extragalactic sources. Their diameters are too small, by factors of 4--10, for them to be seen through the scattering region in front of \\sgra. However, they can be used, in conjunction with the heavily scattered masers, to set constraints on the angular extent of the region. Our fields show a paucity of sources near \\sgra; a previous survey with more uniform sky coverage, but at a lower sensitivity also shows a paucity. Such a deficit could arise if the scattering toward the GC is so severe that our (and previous) observations resolve out extragalactic sources. The sources reported here are combined with angular broadening measurements of \\sgra\\ and OH masers and free-free emission and absorption measurements from the literature. These data are then used in a likelihood analysis to determine the model parameters of the GC scattering region (\\cite{lc98})." }, "9804/astro-ph9804010_arXiv.txt": { "abstract": "In this paper, the third of a series dedicated to the investigation of the nuclear properties of spiral galaxies, we have {\\it (i)} modelled the {\\tt WFPC2} F606W nuclear surface brightness profiles of 41 spiral galaxies presented in Carollo et al.\\ 1997c, 1998 with the analytical law introduced by Lauer et al.\\ 1995, and {\\it (ii)} deprojected these surface brightness profiles and their analytical fits, so as to estimate the nuclear stellar densities of bulges of spiral galaxies. We find that the nuclear stellar cusps (quantified by the average logarithmic slope of the surface brightness profiles within 0.1$''$-0.5$''$) are significantly different for $R^{1/4}$-law and exponential bulges. The former have nuclear properties similar to those of early-type galaxies, i.e. similar values of nuclear cusps for comparable luminosities, and increasingly steeper stellar cusps with decreasing luminosity. By contrast, exponential bulges have (underlying the light contribution from photometrically distinct, central compact sources) comparative shallower stellar cusps, and likely lower nuclear densities, than $R^{1/4}$-law bulges. ", "introduction": "The galactic nuclei are the repositories of low angular momentum material sunk to the centers over the lifetime of the parent systems. Therefore, they are likely to hold answers to important questions related with the origin of structure in the parent galaxies. In this perspective, establishing the demographics of galactic nuclei along the entire Hubble sequence lies at the heart of our understanding of the complex process of galaxy formation and evolution. Observations of nearby ellipticals and lenticulars with the Faint Object Camera ({\\tt FOC}), the Wide Field Planetary Camera ({\\tt WF/PC}) and the Wide Field Planetary Camera-2 ({\\tt WFPC2}) aboard the {\\it Hubble Space Telescope} (HST) have revealed that the nuclei of these galaxies are complex environments (e.g., Crane et al.\\ 1993; Jaffe et al.\\ 1994; Lauer et al.\\ 1995, hereafter L95; Forbes et al.\\ 1995; Carollo et al.\\ 1997a, hereafter C97a; Carollo et al.\\ 1997b, hereafter C97b; Faber et al.\\ 1997, hereafter F97). They show surface brightness profiles that increase down to the innermost point measurable at HST resolution, i.e., $I(r) \\propto r^{-\\gamma}$ as $r \\rightarrow 0$ (where $I(r)$ is the surface brightness at the radius $r$, and $\\gamma > 0$); furthermore, several galaxies host stellar and gaseous disks, unresolved nuclear spikes, double nuclei. These inner features might possibly be related to the presence of massive black holes (e.g., Lauer et al.\\ 1996). By contrast, much is still to be learned about the nuclear properties of nearby spiral galaxies at HST resolution. F97 found that the surface brightness profiles of the three Sa-Sb bulges present in their sample show a behaviour similar to that of early-type spheroidals of comparable luminosity. The same result was found by Phillips et al.\\ (1996) for the three spirals of type earlier than Sc contained in their {\\tt WF/PC} F555W sample of 20 disk galaxies. Furthermore, Phillips et al.\\ found that later type spirals show instead (almost) flat nuclear profiles, and suggested that the nuclear properties of disk galaxies are more closely related to those of nucleated dwarf galaxies than to those of elliptical galaxies. Further exploration is necessary to assess how the nuclear properties scale with the properties of the spheroidal component. This is likely to provide feedback information about the epoch and processes of nucleus, bulge, and, ultimately, galaxy formation. In order to address these issues, we have performed a {\\tt WFPC2} snapshot survey in the F606W filter of the nuclei of 107 (mostly Sa to Sc) disk galaxies. In paper I (Carollo et al.\\ 1997c) and paper II (Carollo et al.\\ 1998) we have presented the 75 targets imaged so far within our program. Our analysis shows that bulge-like structures are present in most of the galaxies. While in some cases these are ``classical'', smooth, featureless $R^{1/4}$-law bulges, in others they are better fitted by an exponential profile (see also Courteau, de Jong \\& Broeils 1996, and references therein, for similar results). The exponential bulges include two classes of objects: {\\it (i)} dwarf-looking systems, whose surface brightness profiles within $\\approx 15''$ are well fitted by a single exponential. These galaxies are strongly bulge-dominated; their surrounding, faint regions (``disk/halo'') show no signs of spiral arms, and have typically a quiescent, i.e. non star forming, appearance. {\\it (ii)} Small exponential bulges embedded in dominant (spiral-armed/star-forming) surrounding disks, i.e., the inner exponential structures of double-exponential fits to the surface brightness profiles within $\\approx 15''$. The exponential bulges as-a-class are statistically fainter than the featurless, smooth $R^{1/4}$ bulges, for constant disk luminosity and Hubble type. Resolved, central compact sources are found in most of the exponential bulges; the hosts of central compact sources often contain a barred structure. The nature of these compact source, and in particular their relation with e.g., star clusters and Seyfert 2 nuclei, is discussed in Carollo (1998). In this paper, the third of the series, we investigate the relation between the nuclear structure of spiral galaxies and the physical properties of inner disks and/or bulges. In particular, we {\\it (i)} present the results of modeling the nuclear surface brightness profiles with the analytical law introduced by L95 (for the 43 galaxies of paper I and II for which we could perform the measurements), {\\it (ii)} deconvolve the surface brightness profiles and their analytical fits in order to estimate the nuclear stellar densities, {\\it (iii)} study the nuclear properties as a function of the global properties discussed in papers I and II (e.g., $R^{1/4}$ against exponential bulges), and {\\it (iv)} compare the nuclear properties of our sample with those observed in early-type galaxies. The paper is organized as follows. In section 2 we briefly summarize the properties of the sample, the data used in our investigation, the procedure adopted for the data reduction, and the steps performed to derive the surface brightness profiles. In section 3 we present the results of the analytical fits applied to the surface brightness profiles, and of deconvolving data and models in spherical symmetry. In section 4 we investigate the dependence of the nuclear properties on global galactic properties. We conclude in section 5. ", "conclusions": "In this paper we have investigated the relation between the nuclear structure of spiral galaxies and the physical properties of their bulges. In particular, we have {\\it (i)} modelled the {\\tt WFPC2} F606W nuclear surface brightness profiles of 41 spiral galaxies with the analytical law introduced by L95 (data from papers I and II), and {\\it (ii)} deconvolved the surface brightness profiles and their analytical fits in order to estimate the nuclear stellar densities of disk galaxies. Our main result is that $R^{1/4}$-law bulges and exponential bulges have significantly different nuclear stellar cusps and densities. Specifically, $R^{1/4}$-law bulges have steep stellar cusps which steepen with decreasing luminosity; furthermore, their stellar cusp slopes and densities are similar in values to those of early-type systems of comparable luminosity. By contrast, in exponential bulges, the inward extrapolations underlying the light from the compact sources which sit in their very centers imply rather shallow stellar cusps and, very likely, relatively low nuclear stellar densities. \\bigskip \\bigskip \\noindent {\\bf Acknowledgements} We heartly thank Tim Heckman and Colin Norman for helpful discussions, and the anonymous referee for constructive comments to an earlier version of this paper. CMC is supported by NASA through the grant HF-1079.01-96a awarded by the Space Telescope Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., for NASA under contract NAS 5-26555. This research has been partially funded by grant GO-06359.01-95A awarded by STScI, and has made use of the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, Caltech, under contract with NASA. \\bigskip \\bigskip \\noindent \\begin{center}{{\\bf Appendix A.} Details on the Analytical Fits}\\end{center} \\medskip We used the same fitting procedure described in C97a; we addess to this reference for further information. This was carried out in two steps: (1) The first step isolated primary from secondary minima of $\\chi^2$. The $\\chi^2$ values were computed on a grid of points uniformly distributed on a wide hypercube in parameter space (with dimension equal to the number of free parameters). Once a minimum value was found, a new, smaller, hypercube was placed on that location, and the procedure iterated. (2) The minimum value found on the hypercube was then used as starting point to initialize a downhill simplex minimization. We tested the procedure on simulated data, and verified that it recovered the initial values with high accuracy. We accepted as final the fits associated with the absolute minimum of $\\chi^2$. Deconvolutions of {\\tt WF/PC} data have been proven to be very reliable (e.g., L95). However, in our analysis, we chose not to apply any deconvolution to the post-refurbishment {\\tt WFPC2} images, and to correct for PSF-blurring while modeling the light profiles. Therefore, the models were convolved with the appropriate PSFs before being compared to the data. Since pointlike sources with adequate S/N located near to the nuclei were not available for most of the galaxies, we computed the PSFs by running Tinytim (Krist 1992). Focus drifts and breathing modify the PSF profile and affect the flux within a 1 PC pixel radius up to 10\\% (and within 5 PC pixels up to 5\\%; Suchkov \\& Casertano 1997). Therefore, the simulated PSFs obtained by construction at the nominal focus position are in principle of similar quality than PSFs derived from archival stars. Furthermore, our approach of convolving the models rather than deconvolving the data minimizes the effects of using a possibly non-perfect PSF. \\bigskip \\bigskip \\noindent \\begin{center}{{\\bf Appendix B.} Classifying $R^{1/4}$-law or exponential bulges outside $1''$}\\end{center} \\medskip In order to ensure that the classification of a bulge as an $R^{1/4}$-law or an exponential structure is valid on a radial range entirely different from that used in the derivation of $\\langle \\gamma \\rangle$ (equal to 0.1$''$-0.5$''$), we performed as a test the $R^{1/4}$-law and exponential fits after excluding the data inside the innermost $1''$. This value is a compromise between a radius large enough to exclude the range where $\\langle \\gamma \\rangle$ is computed, and small enough to still allow the detection of the small, disk-embedded exponential bulges (pentagons in the figures). As an example, the results of the test are illustrated in Figure 8 for the same four galaxies presented in Figure 6. The solid lines represent either a single exponential (left panels) or a double exponential (right panels) fit; the dashed lines represent either a single $R^{1/4}$-law (left panels) or an $R^{1/4}$-law plus exponential (right panels) fit. Two different scales are used for the abscissa for the galaxies in the left and right panels, consistently with the different scales of their bulge components. An offset of two magnitudes has been applied to ESO482G17 for plotting purposes. There are two important points to note: {\\it (i)} the kind of profile which provides the bulge classification given in papers I and II, i.e. exponential or $R^{1/4}$-law, still provides a better fit to the inner galactic regions, even when the innermost $1''$ is excluded from the fits; {\\it (ii)} the alternative profile with respect to the one that provides the classification given in papers I and II generally provides (not only a worse fit but also) physically meaningless best fit parameters (e.g., for ESO482G17, the $R^{1/4}$-law best fit of Figure 8 has an $R_e\\sim385''$). We conclude that the distinction between $R^{1/4}$-law and exponential bulges holds in a radial range which excludes the one used to derive $\\langle \\gamma \\rangle$, and that the difference in nuclear cusp slopes $\\langle \\gamma \\rangle$ between $R^{1/4}$-law and exponential bulges has a physical origin. We retained the bulge parameters presented in papers I and II for our discussion, since those fits gave an overall better description of the profiles. \\bigskip \\bigskip" }, "9804/astro-ph9804226_arXiv.txt": { "abstract": "An object discovered during an infrared survey of the field near the quasar B2 0149$+$33 has an emission line at 2.25\\,$\\mu$m that we interpret as H$\\alpha$ at a redshift of 2.43. The K-band image shows two compact components 10\\,kpc apart surrounded by more extended emission over $\\sim 20$\\,kpc. The H$\\alpha$ emission appears to be extended over $\\sim 15$\\,kpc (2$^{\\prime\\prime}$) in a coarsely sampled (0\\farcs8/pixel) image. The star formation rate may be as high as 250 -- 1000\\,M$_\\odot$\\,yr$^{-1}$, depending on the extinction. Alternatively, the line may be powered by an active nucleus, although the probability of serendipitously discovering an AGN in the survey volume is only $\\sim 0.02$. The increasing number of similar objects reported in the literature indicate that they may be an important, unstudied population in the high redshift universe. ", "introduction": "Discovering the properties of galaxies at redshifts greater than one requires techniques that can readily distinguish the high redshift objects from those at lower redshift that predominate in any deep image. Multi-wavelength approaches are especially fruitful and have been dominated to date by optical and radio surveys. The use of photometric redshifts based on the strong Lyman break redshifted into the optical band has been one of the most successful of these techniques (Steidel et al. 1996, and references therein). Methods using infrared images are beginning to uncover objects not easily discovered with optical or radio methods that may, nevertheless, constitute a significant fraction of the high redshift population. These include objects distinguished by unusually red colors (Elston, Rieke, \\& Rieke 1988; Soifer et al. 1994; Graham et al. 1994; Hu \\& Ridgway 1994; Cowie et al. 1994; Dey, Spinrad, \\& Dickinson 1995) and objects with emission lines redshifted to infrared wavelengths (Songaila et al. 1994; Thompson, Mannucci, \\& Beckwith 1996, hereafter TMB96; Malkan, Teplitz, \\& McLean 1995; Bechtold et al. 1997). These objects have been interpreted as elliptical galaxies (Graham et al. 1994; Hu \\& Ridgway 1994; Dunlop et al. 1996), young ``protogalaxies'' undergoing bursts of star formation (Eisenhardt \\& Dickinson 1992; Graham \\& Dey 1996; Malkan, Teplitz, \\& McLean 1996; Yee et al. 1996; Bechtold et al. 1997), or active galactic nuclei (Cowie et al. 1994; Dey, Spinrad, \\& Dickinson 1995). All of these populations could be significant for cosmology, since the first case implies massive galaxy formation at redshifts greater than $\\sim$3, the second indicates a substantial population of young galaxies that can be discovered only in infrared surveys, and the third implies a population of infrared bright active galactic nuclei (AGN) comparable in number density to the populations of AGN discovered by more traditional methods. Only a few such objects have been studied in enough detail to reveal redshifts, source morphologies, and colors. In the course of our survey for emission line galaxies (TMB96), an object was discovered near the quasar B2 0149$+$33 that had an emission line at 2.25\\,$\\mu$m, the same wavelength as the quasar's H$\\alpha$ line at a redshift of 2.43; we call this object TMB 0149 - cK39, or simply cK39. A spectrum between 1.5 and 2.4 $\\mu$m, presented and discussed here, confirmed the presence of an emission line nearly coincident in wavelength with that of the quasar. The object is very red, making it unusual among known, distant galaxies. This paper describes the results of these observations and suggests that such objects may be common but previously unobservable owing to the lack of optical emission. ", "conclusions": "The emission line object, TMB 0149 $-$ cK39, appears to be a pair of galaxies undergoing a merger at redshift of 2.4. If cK39 derives a substantial part of its luminosity from star formation, the formation rate is as high as 1000 M$_\\odot$ yr$^{-1}$, an exceptionally large value that is rarely seen in other starforming galaxies. On the other hand, the system could contain at least one active nucleus, perhaps with some contribution from star formation. The individual components appear to be 6 and 7 kpc in extent with the centers separated by 10 kpc, consistent with a merger-induced fuelling of the nuclear activity. An extinction of A$_{\\rm V} \\sim 1.7^{\\rm m}$ is required to produce the observed R-K color of 5.5, requiring the presence of a significant amount of dust in order to suppress the ultraviolet light and redden the galaxies. Although gravitational lensing by a foreground galaxy or cluster of galaxies could enhance the brightness of the emission, there is little evidence of such a galaxy or cluster along the line of sight. The growing number of very red galaxies at high redshift indicates that there are new populations uncovered only in infrared surveys (e.g. Graham \\& Dey 1996; Malkan, Teplitz, \\& McLean 1996; and references therein). In a separate paper (Thompson et al. in preparation), we present statistics that show extremely red objects (R-K$^\\prime > 6$) have a sky density of order 500 deg$^{-2}$ for K$^\\prime$ $\\leq$19.75, so cK39 may, indeed, be part of a larger population. These results underscore the importance of employing a number of different techniques for exploring the epoch of early star formation and demonstrate that a better understanding of these objects is important to an understanding of galaxy formation in the early universe." }, "9804/astro-ph9804295_arXiv.txt": { "abstract": "s{Direct and indirect detection rates of relic neutralinos are reviewed in the framework of the Minimal Supersymmetric Standard Model. The theoretical estimates are compared with the most recent experimental limits from low--background detectors and neutrino telescopes. The properties of neutralino under the hypothesis that preliminary experimental results of the DAMA/NaI Collaboration may be indicative of a yearly modulation effect are examined.} \\normalsize\\baselineskip=15pt \\vspace{-5pt} ", "introduction": "\\vspace{-5pt} Supersymmetric theories predict a large number of particles in excess to the Standard Model ones. If the R--parity is conserved, the lightest among all the supersymmetric particles (LSP) must be stable. This feature makes the LSP a dark matter candidate, since this particle can be present today as a relic from the early stages of the evolution of the Universe. Different candidates have been proposed in the framework of supersymmetric theories: the neutralino or the sneutrino\\cite{sneutrino} in gravity mediated models, the gravitino\\cite{gravitino} or some messenger fields in gauge mediated theories\\cite{messenger}, the axino\\cite{axino}, stable non--topological solitons (Q--balls)\\cite{Qballs} or others. In this paper we will focus on the most promising among all the different candidates, the neutralino, which is defined as the lowest mass linear superposition of photino ($\\tilde \\gamma$), zino ($\\tilde Z$) and the two higgsino fields ($\\tilde H_1^{\\circ}$, $\\tilde H_2^{\\circ}$), i.e. $\\chi \\equiv a_1 \\tilde \\gamma + a_2 \\tilde Z + a_3 \\tilde H_1^{\\circ} + a_4 \\tilde H_2^{\\circ}$. The aim of this review is at providing the latest results on the calculation of different kinds of detection rates of relic neutralinos, in the framework of the Minimal Supersymmetric extension of the Standard Model (MSSM), constrained by the most recent experimental data coming from accelerator physics. We do not discuss here the details of the model, for which we refer to Refs.\\cite{pinning,extending} and to the references quoted therein. We only recall the standard assumptions employed here: i) all trilinear parameters are set to zero except those of the third family, which are unified to a common value $A$; ii) all squarks and sleptons soft--mass parameters are taken as degenerate: $m_{\\tilde l_i} = m_{\\tilde q_i} \\equiv m_0$; iii) the gaugino masses are assumed to unify at $M_{GUT}$, and this implies $M_1= (5/3) \\tan^2 \\theta_W M_2$ at the electroweak scale. After these conditions are applied, the free parameters are: $M_2, \\mu, \\tan\\beta, m_A, m_0, A$. The parameters are varied in the following ranges: $10\\;\\mbox{GeV} \\leq M_2 \\leq 500\\;\\mbox{GeV},\\; 10\\;\\mbox{GeV} \\leq |\\mu| \\leq 500\\;\\mbox{GeV},\\; 65\\;\\mbox{GeV} \\leq m_A \\leq 500\\;\\mbox{GeV},\\; 100\\;\\mbox{GeV} \\leq m_0 \\leq 500\\;\\mbox{GeV},\\; -3 \\leq {\\rm A} \\leq +3,\\; 1.01 \\leq \\tan \\beta \\leq 50$. In our analysis the supersymmetric parameter space is constrained by all the experimental limits on Higgs, neutralino, chargino and sfermion searches at accelerators. Moreover, the constraints due to the $b \\rightarrow s + \\gamma$ process\\cite{LEP} are satisfied. In addition to the experimental limits, we require that the neutralino is the lightest supersymmetric particle. Finally, the regions of the parameter space where the neutralino relic abundance exceeds the cosmological bound, i.e. $\\Omega_{\\chi}h^2 > 1$, are also excluded. \\vspace{-7pt} ", "conclusions": "\\vspace{-10pt} In this paper we have reported the most recent calculations of the direct and indirect detection rates of relic neutralinos in the framework of the Minimal Supersymmetric Standard Model. We have shown that the theoretical estimates of the detection rates may be at the level of the present experimental sensitivities of low--background detectors and neutrino telescopes. For many supersymmetric configurations, and for median values of the astrophysical parameters which enter in the calculations of the detection rates, the predicted signals may already exceed the present experimental bounds. This shows that the different experimental efforts to search for relic particles are potentially able to deeply investigate the possibility that neutralino is a component of the dark matter of the Universe. An interesting preliminary analysis of the DAMA/NaI Collaboration has shown that their data are compatible, at 90\\% C.L., with a modulation signal of the direct detection rate. The features of a neutralino able to satisfy the prerequisites of this signal have been analyzed and it has been shown that many configurations are compatible with a dark matter scenario where the neutralino is the major component, both on galactic and cosmological scales. However, we have to remind here that the occurrence of a possible modulation effect will necessarily require further investigations with much higher statistics. This project is currently under way. \\vspace{-7pt}" }, "9804/astro-ph9804341_arXiv.txt": { "abstract": "Primordial Black Holes (PBH) may have formed from the collapse of cosmic string loops. The spectral shape of the PBH mass spectrum can be determined by the scaling argument for string networks. Limits on the spectral amplitude derived from extragalactic $\\gamma$-ray and galactic $\\gamma$-ray and cosmic ray flux observations as well as constraints from the possible formation of stable black holes remnants are reanalyzed. The new constraints are remarkably close to those derived from the normalization of the cosmic string model to the cosmic microwave background anisotropies. ", "introduction": "Cosmic strings (CS) are linear topological defects that are believed to originate during phase transitions in the very early Universe \\cite{VSrev,HKrev,RBrev}. Here, we consider the ``standard\" CS model \\cite{Zel,V81}, according to which the network of linear defects quickly reaches a ``scaling\" solution characterized by having the statistical properties of the string distribution independent of time if all lengths are scaled to the Hubble radius ($R_H = c t$, where $c$ is the speed of light). Cosmic string loops (CSL) are continually formed by the intersection and self-intersection of long CS (infinite CS or CSL with radius of curvature larger than $R_H$). After formation, a loop oscillates due its own tension and slowly decays by emitting gravitational radiation. The initial length of a CSL is $l(t) = \\alpha R_H$, where $\\alpha$ is expected to be $\\sim G \\mu/c^2$. The mass of a CSL is $m(t) = l(t) \\mu$, where $\\mu$ is the mass per unit of length of the string Since CS also lead to cosmic microwave background (CMB) anisotropies, the string model can be normalized by the recent COBE observations giving the constraint \\cite{LP93,ACSSV} \\be \\label{cmbnorm} G\\mu / c^2 \\leq 1.7(\\pm0.7) \\times 10^{-6} \\ee Our assumption is that a distribution of PBH was formed by the collapse of a fraction $f$ of CSL \\cite{SH,AP}. Hence, from the observational consequences of a present surviving distribution of PBH we can derive updated constraints on the CS scenario \\cite{jru}. These constraints are important because: {\\it i-)} They may indicate new ways to search for direct signatures from CS; {\\it ii-)} They may provide constraints on CS models with symmetry breaking scale $\\mu^{1/2}$ smaller than $10^{16}$ GeV which are not constrained by CMB and large-scale structure data; and {\\it iii-)} They may provide tighter limits than the CMB on CS models with $G \\mu / c^2 \\sim 10^{-6}$. Because CS do not dominate the energy density of the Universe, the CS network must lose energy. We derive the rate of CSL production ${{dn_l} \\over {dt}}$ from the conservation of string energy in the ``scaling\" scenario, \\[ \\label{consmass} {\\dot \\rho_{\\infty}} - 2 H \\rho_{\\infty} = - {{dn_l} \\over {dt}} \\alpha \\mu t \\, , \\] where $\\rho_{\\infty} = \\nu \\mu c^{-3} t^{-2}$ is the energy density in long strings and $\\nu$ is proportional to the average number of long strings crossing each Hubble volume. Hawking \\cite{SH} and Polnarev and Zembowicz \\cite{AP} first postulated that a fraction $f$ of the CSL could collapse within its Schwarzchild radius and then form a BH. More recently, Caldwell and Casper \\cite{CC} have performed numerical simulations to determine $f$ and found \\be \\label{fvalue} f = 10^{4.9 \\pm 0.2} (G \\mu / c^2)^{4.1 \\pm 0.1} \\, . \\ee The BH are sufficiently small that they lose mass due to the Hawking evaporation process. The fraction of the critical density of the Universe in PBH today due to collapsing CSL is (see \\cite{jru} and references quoted therein) \\be \\label{omegaPBH} \\Omega_{PBH}(t_o) = \\frac{1}{\\rho_{crit}(t_o)} \\int_{t_*}^{t_o} dt' \\frac{dn_{BH}}{dt'} m(t',t_o) \\, , \\ee where $t_o$ is the present age of the Universe; $t_*$ is formation time for a PBH with mass $M_* = 4.4 \\times 10^{14} h^{-0.3} \\; \\mbox{g}$, which is expiring today; $m(t',t_0)$ is the mass of a PBH formed at a time $t'$ at a later time $t$; and $h$ is the Hubble parameter in units of $100 \\mbox{km} \\mbox{s}^{-1} \\mbox{Mpc}^{-1}$. PBH formed at times $t < t_* \\; (M < M_*)$ do not contribute to this integral because they will have evaporated by today. If we assume for simplicity that PBH with mass $M > M_*$ will have evaporated little by the present time, we can approximate $m(t',t_0)$ by $\\alpha \\mu c t'$. ", "conclusions": "We have taken advantage of the recent numerical simulations to better understand PBH formation. The observational consequences of a PBH distribution were used to constrain the CS scenario. We have found that the limits on $G \\mu/c^{2}$ are comparable to those stemming from other criteria. Unless the mass of the BH remnants is larger than $10^3 m_{pl}$, these remnants will contribute negligibly to the dark matter of the Universe, even if the BH formation rate has the maximal value allowed by the $\\gamma$-ray flux constraints. A remnant mass of $10^3 m_{pl}$, however, can arise naturally in some models \\cite{CPW} of BH evaporation. In this case, cosmic strings could consistently provide an explanation for the origin of cosmological structure, for the dark matter, and for the origin of the extragalactic $\\gamma$-ray and Galactic cosmic ray backgrounds around $100 MeV$." }, "9804/astro-ph9804207_arXiv.txt": { "abstract": "We present new elements in the identification of the lens-system producing the 4 images of the BAL quasar H1413+117, based on the recent HST/NICMOS-2/F160W observations. After a careful PSF subtraction of the 4 images of the quasar, the residual H$_{F160W}$ image reveals the presence of a faint object ($H\\sim 20.5$) within the region enclosed by the 4 quasar images. This object corresponds to a single galaxy: the primary lens of the lens-system. We also identify the galaxies around the Cloverleaf which had been proposed to belong to a galaxy cluster/group at high redshift (Kneib et al 1998): the other component in the lens-system that provides the additional ``external\" shear. For these galaxies, we have derived a likely redshift based upon their R$_{F702W}$, I$_{F814W}$ and H$_{F160W}$ magnitudes. We find that most of them are consistent with belonging to a galaxy cluster/group with mean redshift $\\overline{z}=0.9 \\pm 0.1$. Furthermore we detect 2 very red objects ($I-H\\sim 4$): the faintest one has no observed optical (R$_{F702W}$ and I$_{F814W}$) counterpart, while the brightest has a predicted redshift around $z\\sim 2$, and may be identified with one of the Cloverleaf absorbers. This gravitational-lens system constitutes an excellent target for IR imaging/spectroscopy with the new generation of 8m ground-based telescopes. ", "introduction": "The excellent quality and broad wavelength coverage of current observations of gravitational lensing systems have allowed to unveil part of their mysteries. The primary lens is often detected. The immediate surrounding of the multiply imaged quasars/galaxies is studied in great detail and generally shows some galaxy clustering ({\\it e.g.} Tonry 1998, Hjorth \\& Kneib 1998) or even sometimes X-ray cluster emissions (Hattori et al 1997, Chartas et al 1998). Furthermore, the measure of a time-delay in Q0957+561/PG1115 (Kundic et al 1997, Schechter et al 1997) has strengthened the interest of a detailed study of these gravitational lens systems in order to use them as a cosmological tool. Since the identification by Magain et al (1988) of the quadrupole lens-system called the Cloverleaf, 4 images of the BAL quasar H1413+117 at z=2.558, many efforts have been dedicated to a direct search of the lens or of elements of the lens-system. Early models of the lens-system have involved one or two galaxy-lenses very close to the line-of-sight toward the quasar (Kayser et al, 1990). A more recent analysis showed that an external shear was needed to model this system correctly (Keeton, Kochaneck \\& Seljak 1997), and indeed it is probably related to the existence of an overdensity of galaxies nearby, as detected by Kneib et al (1998). The lensing geometry, amplification and time-delays are quite sensitive to the parameters of the galaxy-lens expected to be located amid the 4 images of the quasar. A positive detection of the galaxy-lens would bring stringent constraints in the modeling of the lens-system. However, despite relatively deep searches, either in R and I imaging with the HST (Turnshek et al, 1997; Kneib et al, 1998) or K imaging with the Keck telescope (Lawrence et al 1996), the lensing galaxy has not been detected. The galaxy cluster recently revealed near the Cloverleaf (Kneib et al, 1998) was assumed to be at $z\\sim 1.7$ as this corresponds to the mean value of the narrow absorption line-systems observed in the quasar spectra (z=1.44, 1.66, 1.87, 2.07 and 2.09: Turnshek et al, 1988; Magain et al, 1988, Monier et al 1998). Combining the IR image with the R$_{F702W}$ and I$_{F814W}$ WFPC-2 images of this system can allow to estimate the likely redshift for the faint galaxies surrounding the Cloverleaf. The recently acquired HST/NICMOS-2/F160W observation consists of a unique dataset to help answer both the existence of the lensing galaxy and to constrain the distance of the nearby galaxies. The HST/NICMOS-2 data are presented in Sect.2, while in Sect. 3 we discuss the identification of the lensing galaxy after the PSF subtraction of the 4 quasar images. The redshift estimates (derived from photometry) of the surrounding galaxies are explained in Sect.4. The discussion and concluding remarks are provided in Sect.5. Throughout this paper we use $H_0=$50 km/s/Mpc and $\\Omega_0=$1. ", "conclusions": "Two main results have been obtained from the NICMOS-2 data. For the first time the galaxy-lens, H1, close to the line of sight toward the quasar, has been identified. Its position with respect to the quasar line-of-sight is found to be similar (within the uncertainties) to the one derived in the various Cloverleaf gravitational lens models ({\\it e.g.} Kneib et al, 1998). It remains difficult to estimate the redshift of the galaxy-lens H1 because the PSF subtraction leaves an increased background noise in the region amid the 4 quasar images. Yet, a redshift estimate around 1.0 or higher is consistent with the H$_{F160W}$ magnitude and the I$_{F814W}$ lower limit magnitude we have derived for H1. Clearly deep spectroscopic data are needed to solve for the determination of its redshift. We find also that there is a unique galaxy-lens, in contradiction to some early models in which two galaxy-lenses had been envisaged (Kayser et al, 1990). Assuming that H1 is around $z\\sim 1$, and has similar colors and absolute magnitude than the nearby galaxies gives a Mass-to-light ratio of M($<5.1$kpc)/L$_B$ $\\sim$ 25 M/L$_{B\\odot}$. With regard to the Cloverleaf environment, we show that 8 nearby galaxies have a most probable redshift around 0.9, giving credit to the presence of a galaxy cluster/group along the line of sight to the Cloverleaf. In our previous modelling (Kneib et al, 1998), we assumed for this galaxy cluster/group a redshift of 1.7, as a mean of the redshifts of the 4 absorbers silhouetted on the quasar spectrum. This value should be revised. The location of the galaxy-lens is now known from the NICMOS-2 observations and will be implemented in a new model of the lens-system. One of the faint galaxies surrounding the Cloverleaf appears to be at a larger redshift, around 2, and might be related with the absorber at z$=$2.07 or 2.09 (Monier et al 1998). Further IR imaging/spectroscopy of these galaxies should remove the remaining uncertainties of the Cloverleaf lens-system." }, "9804/astro-ph9804031_arXiv.txt": { "abstract": "We present data for 18 blazars observed with the X--ray satellite {\\sl ASCA}, half of which were also observed contemporaneously with the EGRET instrument onboard {\\sl Compton Gamma-ray Observatory} as parts of multi-wavelength campaigns. The observations show a clear difference in the spectra between three subclasses of blazars, namely the High-energy peaked BL Lac objects (HBLs), Low-energy peaked BL Lac objects (LBLs), and quasar-hosted blazars (QHBs). The \\asca X--ray spectra of HBLs are the softest, with the power law energy index $\\alpha \\sim 1 - 2$, and they form the highest observable energy tail of the low energy (LE, synchrotron) component. The X--ray spectra of the QHBs are the hardest ($\\alpha \\sim 0.6$) and are consistent with the lowest observable energy end of the high energy (HE, Compton) component. For LBLs, the X--ray spectra are intermediate. We find that the radiation process responsible for the HE peak for HBLs {\\sl can} be explained solely by Doppler-boosted Synchrotron-Self-Compton (SSC) emission, with the Doppler factor $\\delta$ consistent with the VLBI and/or $\\gamma$--ray variability data. For many QHBs, on the other hand, the $\\gamma$--rays {\\sl cannot} be solely due to the SSC mechanism unless $\\delta$ is significantly in excess of that inferred from VLBI data. We consider an alternative scenario consistent with the measured values of $\\delta$, where the SSC component is still present in QHBs and it dominates in the X--ray band, but it is below the observed $\\gamma$--ray spectrum. With an assumption that the peak of the SSC emission is on the extrapolation of the X--ray spectrum, and adopting $\\delta$ of 10, we infer the magnetic field $B$ to be 0.1 -- 1 Gauss, and Lorentz factors $\\gamma_{b}$ of electrons radiating at the peak of the $\\nu F(\\nu)$ spectrum of $\\sim 10^{3}$ for QHBs; this is much lower than $\\gamma_{b} \\sim 10^{5}$ for HBLs, even though the values of $B$ are comparable in the two sub-classes. This difference of $\\gamma_{b}$ is most likely due to the large photon density expected in QHBs (e.g. from thermal components visible in these objects) as compared with that of HBLs; Compton upscattering of these photons may well provide the observed GeV flux. ", "introduction": "The overall electromagnetic spectra of blazars -- a class of active galactic nuclei (AGNs) that includes BL Lac objects and Optically Violently Variable (OVV) quasars -- are believed to be dominated by Doppler-boosted radiation from relativistic jets pointing closely to our line of sight (\\cite{blandford78}; \\cite{blandford79}; \\cite{urry95} for a review of radio loud AGNs). The VLBI studies of these objects show compact radio cores on milli-arcsecond angular scale with jet-like structures which often show superluminal motion, with apparent speeds $v/c \\sim 5 - 10$ (e.g., \\cite{vermeulen}). Apparent variability time scale and luminosity amplification depend on various powers of the ``beaming'' (Doppler) factor $\\delta$ (e.g., \\cite{lind}), defined via $\\delta = \\Gamma_{j}^{-1} (1 - \\beta \\cos \\theta)^{-1}$, where $\\Gamma_{j}$ is the bulk Lorentz factor of the emitting matter, $\\beta = v/c$, and $\\theta$ is the angle of motion with respect to the line of sight. Blazars are commonly detected as $\\gamma$--ray sources. The EGRET instrument onboard the {\\sl Compton Gamma-Ray Observatory} (\\cgro) has so far detected emission in the GeV range from $\\sim$ 50 blazars (\\cite{fichtel94}; \\cite{thompson95}; \\cite{mattox97}; \\cite{mukherjee97}); $\\gamma$--ray emission has been detected up to the TeV range from the nearby BL Lac objects Mkn~421 and Mkn~501 with ground-based Cherenkov telescopes (\\cite{punch92}; \\cite{petry93}; \\cite{quinn96}; \\cite{bradbury97}). As these sources show large luminosity and compact emission regions in the spectral range where the opacity to pair production via $\\gamma\\gamma\\rightarrow e^{+}e^{-}$ is large, it is generally accepted that the $\\gamma$--ray emission is anisotropic and Doppler-boosted as well (\\cite{maraschi92}; \\cite{mattox93}; \\cite{dondi95}; \\cite{buckley96}), suggesting that the {\\sl entire} observed electromagnetic emission arises in the jet. As these objects emit in practically every observable waveband, any study of the structure and physical conditions in the jets requires broad-band spectral observations, which, given the rapid large amplitude flux variability, must be conducted simultaneously. The overall spectra of blazars have two pronounced components: one peaking at low energies (LE), $10^{13}-10^{17}$ Hz (e.g., \\cite{sambruna96}), and another peaking at high energies (HE), in the $\\gamma$--rays (e.g., \\cite{montigny95}). For the blazars that are hosted in quasars (QHBs), and for BL Lac objects discovered via radio-selection techniques (the so-called ``Low-energy peaked BL Lacs'' or LBLs), the LE component peaks in the infrared. For the the majority of BL Lac objects -- those found as a result of their X--ray emission -- it peaks in the ultraviolet or even in the soft X--rays (\\cite{giommi95}; \\cite{sambruna96}; \\cite{padovani96}; \\cite{fossati97a}), and thus they are named ``High-energy peaked BL Lacs (HBLs)'' (\\cite{padovani95}). The local power-law shape, the smooth connection of the entire radio - to - UV (and, for the HBLs, soft X--ray) continuum, as well as the relatively high level of polarization observed from radio to the UV, imply that the emission from the LE component is most likely produced via the synchrotron process of relativistic particles radiating in magnetic field. This view is strongly supported by spectral variability observed in X--rays in a number of HBLs, where the variability at lower X--ray energies lags behind the more energetic X--rays (\\cite{tashiro92}; \\cite{sembay93}; \\cite{kohmura94}; \\cite{tashiro95}; \\cite{takahashi96}) The HE component, on the other hand, peaks in the $\\gamma$--ray band, in the MeV - to - GeV range, and, in the case of a few HBLs, it extends to the TeV range; it is believed to be produced via Comptonization by the same particles that radiate the LE component. The source of the ``seed'' photons, can either be the synchrotron radiation, internal to the jet -- as in the Synchrotron-Self-Compton (SSC) models (\\cite{rees67}; \\cite{jones74}; \\cite{konigl81}; \\cite{ghisellini85}; \\cite{band85}; \\cite{ghisellini89}; \\cite{maraschi92}; \\cite{bloom96}; \\cite{mastichiadis97}). Alternatively, these can be external to the jet, as in the External Radiation Compton (ERC) models: either the UV accretion disk photons (\\cite{dermer92}; \\cite{dermer97}), or these UV photons reprocessed by the emission line clouds and/or intercloud medium (\\cite{sikora94}; \\cite{blandford95}), or else, IR radiation ambient to the host galaxy (\\cite{sikora94}). The ratio of the power in the HE to the LE components is systematically larger for QHBs than for BL Lac objects (\\cite{maraschi94a}; \\cite{dondi95}; \\cite{sambruna97a}; \\cite{ulrich97}; \\cite{fossati97b}). If we assume that the LE component is due to the synchrotron radiation, its peak frequency is determined by the intensity of magnetic field and the distribution function of electron energies, while the location of the HE peak is related to the distribution functions of electron and target photon energies. The ratio of the luminosity of these components ($L_{HE}/L_{LE}$), in the context of this synchrotron plus Compton model, is expected to reflect the ratio of energy densities of photon and magnetic fields in the jet. This paper reports the X--ray spectra of 18 blazars measured by \\asca in the context of their multi-band emission. The \\asca observations and results are described in \\S2, followed by multi-band analysis and discussion in \\S3. Summary of this paper is presented in \\S4. Throughout this paper we use $H_0$=75 km s$^{-1}$ Mpc$^{-1}$, $q_0$=0.5. ", "conclusions": "As we mentioned previously, the two leading models of the high energy emission in blazars invoke Comptonization, of internal (SSC) or external (ERC) seed photons. In the following analysis, we assume that {\\sl both} SSC and ERC processes may operate in blazars. We then estimate the contribution of the SSC emission in the HE component. In order to calculate the predicted luminosity due to the SSC emission, we assume a simple homogeneous model, in which photons are produced in a region of radius $R$ and with a constant magnetic field $B$. We considered the radiation by a single population of relativistic electrons, with a broken power law distribution of Lorentz factors $\\gamma_{el}$, and a break point at $\\gamma_{b}$ (similar to e.g. \\cite{sambruna96}). We also assume that the radiation spectrum of the LE component peaks at a frequency corresponding to that radiated by the electrons with $\\gamma_{b}$. The peak frequency of the synchrotron component in the observer frame, $\\nu_{sync}$, is then given as, when pitch angle is $\\pi$/2: \\begin{equation} \\nu_{sync} = 1.2 \\times 10^{6} \\gamma_{b}^{2} B \\frac{\\delta}{(1+z)} \\quad {\\rm Hz} \\label{eqn1} \\end{equation} where B is in Gauss. If the electron energy is still in the Thomson regime, ($\\gamma_{el} \\times h\\nu_{sync} << m_ec^2$), the expected peak of the SSC component in the observer frame ($\\nu_{SSC}$) is $\\nu_{SSC} = 4 \\gamma_{b}^{2} \\nu_{sync}/3$. The ratio of the observed luminosity of the SSC component $L_{SSC}$ to the observed synchrotron luminosity $L_{sync}$ is: \\begin{equation} \\frac{L_{SSC}}{L_{sync}}=\\frac{u_{sync}}{u_{B}} \\label{eqn2} \\end{equation} where the $u_{sync}=L_{sync}/(4\\pi R^{2}c\\delta^{4})$ is the rest-frame energy density of the synchrotron photons, and $u_{B}=B^{2}/(8\\pi)$ is the magnetic field energy density. To check the validity of the assumption that the observed HE component is solely due to the SSC emission, we calculated the beaming factor ($\\delta$), which is given from above equations: \\begin{equation} \\delta^2 = 1.6\\times 10^{12} \\frac{L_{sync}}{c R^2} \\left(\\frac{L_{sync}}{L_{SSC}}\\right) \\frac{\\nu_{SSC}^2}{\\nu_{sync}^4} \\frac{1}{(1+z)^2} \\label{eqn3} \\end{equation} where $L$ is in erg~s$^{-1}$, $\\nu$ in Hz, $c$ in cm s$^{-1}$, and $R$ in cm. We estimate $R$ from the shortest observed variability (doubling) time scale $\\Delta t$ observed in any wavelength, as given in Table 2. Assuming that $R \\lesssim c \\delta \\Delta t / (1+z)$, then Eq. 3 can be rewritten as: \\begin{equation} \\delta^4 \\gtrsim 1.6\\times 10^{12} \\frac{L_{sync}}{c^3 \\Delta t^2} \\left(\\frac{L_{sync}}{L_{SSC}}\\right) \\frac{\\nu_{SSC}^2}{\\nu_{sync}^4} \\label{eqn4} \\end{equation} where $\\Delta t$ is in s, and other quantities are as in Eq. 3. An application of this equation to the data in Table 2 assuming $L_{sync}=L_{LE}$, $L_{SSC}=L_{HE}$, $\\nu_{sync}=\\nu_{LE}$, $\\nu_{SSC}=\\nu_{HE}$ implies that the lower limits of $\\delta$ for HBLs are $\\sim $3 or less, which is consistent with the VLBI results (cf. \\cite{vermeulen}), and the limits obtained from the arguments of the $\\gamma$--ray opacity (cf. \\cite{dondi95}). However, for 4 QHBs, where the $\\gamma$--ray flux severely dominates the radiative output, we derive values of $\\delta$ that are much larger than the VLBI results (see Table 2). This suggests that an additional emission mechanism -- such as the ERC process -- may contribute significantly in the $\\gamma$--ray regime, dominating over the SSC flux, and the values of $\\nu_{SSC}$ and $L_{SSC}$ are {\\sl very} different than $\\nu_{HE}$ and $L_{HE}$, with the SSC component ``hidden'' well below the ERC component. However, the fact that the QHBs have X--ray spectra which are hard, with $\\alpha \\sim 0.6$, and which are {\\sl not} located on the extrapolation of the synchrotron optical / UV spectra, implies that the X--rays observed in QHBs are due to a separate emission process than synchrotron. The fact that for most of QHBs the $\\gamma$--ray spectra are above the extrapolation of X--ray spectra (Fig. 2) suggests that the dominant process is different for X--rays than it is for $\\gamma$--rays. One explanation is that the SSC process dominates in the X--ray range, while the ERC process dominates in $\\gamma$--rays (\\cite{inoue}). With the assumption that SSC process is dominant in X--rays for QHBs, we estimate the location of the ($\\nu_{SSC}$, $L_{SSC}$) point in the log($\\nu$) -- log($\\nu F (\\nu)$) space by the following method. We assume that it lays on or below the extrapolation of the {\\sl ASCA} spectrum (line (a) in Fig. 3), but above the highest value of $\\nu F (\\nu)$ measured by {\\sl ASCA}. Since the spectra of QHBs generally have $\\alpha < 1$ and thus $\\nu F (\\nu)$ is the highest at the end of the \\asca bandpass (10 keV $\\simeq$ 2$\\times$10$^{18}$Hz), this second limit is equivalent to $\\nu_{SSC} L_{SSC} > 2\\times 10^{18}$ Hz $L_{10 \\rm keV}$ (line (b) in Fig. 3). We further constrain $L_{SSC}$ using Eq. 3; once we assume a given $\\delta$, there is a unique relationship between $L_{SSC}$ and $\\nu_{SSC}$ described as: \\begin{equation} L_{SSC} = 1.6 \\times 10^{12} \\left(\\frac{L_{sync}^{2}}{cR^{2}\\nu_{sync}^{4} \\delta^{2} (1+z)^2} \\right) \\nu_{SSC}^{2} \\quad {\\rm erg~s^{-1}} \\label{eqn5} \\end{equation} where $L$, $R$, $c$, and $\\nu$ are in the same units as in Eqs 3 \\& 4. The VLBI data and $\\gamma$-ray opacity argument suggest that $5 < \\delta < 20$ for most blazars (e.g.,\\cite{vermeulen}; \\cite{dondi95}). This corresponds to the lines (c) and (d) in Fig. 3, respectively for $\\delta$=5 and 20. The above four constraints correspond to the shaded area of Fig. 3, where for illustration, we use the overall spectral energy distribution for the QHB CTA~102. Since we have to use a unique value in calculating the physical parameters, we use $\\delta$ = 10 as a geometrical mean between 5 and 20. The $L_{SSC}$ - $\\nu_{SSC}$ line calculated from Eq. 5 corresponding to $\\delta$ = 10 intersects both the extrapolation of the {\\sl ASCA} spectrum and the highest {\\sl ASCA} value, and the intersections yield the lower and upper values for both $L_{SSC}$ and $\\nu_{SSC}$. We adopt a mean of these values, which are given in Table 2, and plotted in Figure 4c. We used \\ginga data instead of \\asca data for 3C~279 because a simultaneous campaign from radio to $\\gamma$--ray bands was conducted during \\ginga observation. The values for the other blazars where $\\delta$ derived from Eq. \\ref{eqn4} is $<20$, are calculated by assuming $L_{SSC}$ = $L_{HE}$. For LBL AO0235+164 where $\\delta$ derived from Eq. 4 is $>20$, we assume $L_{SSC}$ = $L_{LE}$ because the \\asca spectra of AO0235+164 is thought to be mixture of the LE and HE component, as discussed by \\cite{madejski96} based on the \\rosat and \\asca spectra, so that the above method may be inappropriate. For two QHBs (3C~273, PKS~0208-512) the lower limits of $\\delta$ are $\\sim$5. The fact that observed $\\gamma$-ray flux of PKS~0208-512 is much higher than the extrapolation of \\asca spectrum implies the $\\gamma$-ray peak is not solely due to SSC mechanism. Therefore we applied the above method to this source. On the other hand, since the $\\gamma$-ray spectrum of 3C~273 is below the extrapolation of \\asca spectrum, the $\\gamma$-ray emission is assumed to be due to SSC mechanism so there, we assume $L_{SSC}$ = $L_{HE}$. It is important to note, however, that 3C~273 is unique as compared to other blazars considered here in that the ``blue bump'' is very pronounced, and thus it is unlikely that the jet dominates the entire electromagnetic emission, and therefore, a more complex analysis is necessary (see, e.g., \\cite{montigny97} for further discussion). Once we obtain $L_{SSC}$, and $\\nu_{SSC}$, we can calculate the strength of the magnetic field and the electron Lorentz factor $\\gamma_{b}$ from Eq. (\\ref{eqn1}), (\\ref{eqn2}) and those are given as follows: \\begin{equation} B = 0.27 \\left(\\frac{R_{\\mbox{pc}}}{10^{-2}}\\right)^{-1} \\left(\\frac{\\delta}{10}\\right)^{-2} \\sqrt{\\left(\\frac{L_{sync}}{10^{46}}\\right) \\left(\\frac{L_{sync}}{L_{SSC}}\\right)} \\quad {\\rm Gauss} \\end{equation} \\begin{equation} \\gamma_{b} = 1.8\\times 10^3 \\left(\\frac{R_{\\mbox{pc}}}{10^{-2}}\\right)^{1/2} \\left(\\frac{\\delta}{10}\\right)^{1/2} \\left(\\frac{\\nu_{sync}(1+z)}{10^{13}}\\right)^{1/2} \\left[\\left(\\frac{L_{sync}}{10^{46}}\\right) \\left(\\frac{L_{sync}}{L_{ssc}}\\right)\\right]^{-1/4} \\end{equation} where $R_{\\mbox{pc}}$ is size of emission region in parsecs, and other quantities are as in Eqs 3, 4, \\& 5. As before, the upper limit of the size $R$ can be estimated from the observed time variability ($\\Delta t$) from Table 2, given by $R \\lesssim c\\Delta t \\delta / (1+z)$. Our calculated values of $B$ and $\\gamma_{b}$ are plotted respectively in Figures 4d and 4e. In these Figures, we also plot the values calculated with $R = 0.01$ pc, which would correspond to an observed variability time scale of $\\sim 1$ day. From our analysis, the magnetic field for blazars observed with \\asca is inferred to be 0.1 -- 1 Gauss. The value of $B$ is comparable between the different subclasses of blazars, although $B$ is somewhat lower in HBLs than in QHBs. With these values of $B$, we estimate $\\gamma_{b}$ to be $10^{3} - 10^{4}$ for QHBs, and $10^{5}$ for HBLs. The differences of $\\gamma_{b}$ between different sub-classes of blazars imply that the relativistic electrons are accelerated to higher energies in HBLs than in QHBs. Alternatively, higher $\\gamma_{b}$ in HBLs might be obtained by increasing $\\delta$. However, in those objects, we believe that there is no contribution to the $\\gamma$--ray production from other mechanisms besides SSC, and thus the observed $L_{HE}$ is $L_{SSC}$. In such case, $\\gamma_{b}$ depends on $\\delta$ only linearily (cf. Eq. 7 and $R\\lesssim c\\Delta t \\delta/(1+z)$), and thus varying $\\delta$ to be 5 or 20 respectively decreases or increases our derived $\\gamma_{b}$ only by a factor of two, which is small when compared to the large difference of $\\gamma_{b}$ calculated by us (cf. Fig. 4e). In QHBs the strong optical and UV line emission implies a presence of dense external radiation fields. This means that in the frame of reference of the jet, these can easily dominate over the internal synchrotron radiation, resulting in the ERC emission dominating over the SSC emission in $\\gamma$--ray band (e.g., \\cite{sikora97}). It is likely that the difference of $\\gamma_{b}$ is most likely due to the large photon density in QHBs as compared with that of HBLs. It should be noted that TeV photons have been observed only from HBLs, where we calculate higher values of $\\gamma_{b}$." }, "9804/astro-ph9804096_arXiv.txt": { "abstract": "In this paper we present a new method that can be used for analysis of time of arrival of a pulsar pulses (TOAs). It is designated especially to detect quasi-periodic variations of TOAs. We apply our method to timing observations of PSR B1257+12 and demonstrate that using it it is possible to detect not only first harmonics of a periodic variations, but also the presence of a resonance effect. The resonance effect detected, independently of its physical origin, can appear only when there is a non-linear interaction between two periodic modes. The explanation of TOAs variations as an effect of the existence of planets is, till now, the only known and well justified. In this context, the existence of the resonance frequency in TOAs is the most significant signature of the gravitational interaction of planets. ", "introduction": "The first extra-solar planetary system was discovered by \\cite{Wolszczan:92::} around a millisecond radio pulsar, PSR B1257+12. The three planets orbiting the pulsar have been indirectly deduced from the analysis of quasi-periodic changes in the times of arrival (TOAs) of pulses caused by the pulsar's reflex motion around the center of mass of the system. In the analyses of this kind, it is particularly important to establish a reliable method of distinguishing planetary signatures from possible TOA variations of physically different origin. In the case of PSR B1257+12, it was possible to make this distinction and confirm the pulsar planets through the detection of mutual gravitational perturbations between planets B and C \\cite[]{Wolszczan:94::}, following predictions of the existence of this effect by \\cite{Rasio:92::}, \\cite{Malhotra:92::} (see also \\cite{Malhotra:93::}, \\cite{Rasio:93::} and \\cite{Peale:93::}). Practical methods of detection of the TOA variations caused by orbiting planets include direct fits of Keplerian orbits to the TOA or TOA residual data \\cite[e.g.]{Thorsett:92::,Lazio:95::} and model--independent frequency domain approaches based on Fourier transform techniques \\cite[]{Konacki:96::,Bell:97::}. In fact, it appears that it is best to search for periodicities in TOAs (or post-fit TOA residuals left over from fits of the standard timing models) by examining periodograms of the data, and then refine the search by fitting orbits in the time domain using initial orbital parameters derived from a frequency domain analysis. The presence of planets around a pulsar causes pulse TOA variations which, for planets moving in orbits with small eccentricities, have a quasi-periodic character and generate predictable, orbital element-dependent features in the spectra of TOA residuals. This has led \\cite[]{Konacki:96::} to devising a method of TOA residual analysis based on the idea of a successive elimination of periodic terms applied by Laskar (1992) in his frequency analysis of chaos in dynamical systems. The frequency analysis provides an efficient way to decompose a signal representing the TOA residual variations into its harmonic components and study them in an entirely model-independent manner. As shown in \\cite[]{Konacki:96::} this method works perfectly well under idealized conditions in which covariances among different parameters of the timing model are negligible. In this paper, we present an improved scheme for the frequency analysis of pulsar timing observations in which a successive elimination of periodicities in TOAs is incorporated in the modelling process rather than being applied to the post-fit residuals. This makes the results obtained with our method less sensitive to the effect of significant covariances which may exist between various timing model parameters. We apply the frequency analysis to TOA measurements of the planet pulsar, PSR B1257+12, using a computer code developed to fit spectral timing models to data. We show that our method allows an easy detection of the fundamental orbital frequencies of planets A, B and C in the pulsar system and the first harmonics of the frequencies of planets B and C generated by eccentricities of the planetary orbits. Furthermore, by detecting the effect of perturbations between planets B and C, we demonstrate that the frequency analysis method represents a sensitive, model-independent tool to analyze nonlinear interactions between periodic modes of processes of various physical origins. ", "conclusions": "" }, "9804/astro-ph9804269_arXiv.txt": { "abstract": "We have developed in detail the theory of X-ray line and continuum production due to atomic interactions of accelerated ions, incorporating in our calculations information from a broad range of laboratory measurements. We applied our calculations to the Orion region from which nuclear gamma-ray lines were observed with the COMPTEL instrument on {\\it CGRO}. The accelerated particles which produce this gamma-ray emission via nuclear reactions also produce X-ray lines via atomic interactions. We predict strong line emission in the range from 0.5 to 1 keV, mainly due to de-excitations in fast O ions. While much of the diffuse X-ray emission observed with ROSAT from Orion could be due to accelerated ions, the current X-ray data do not provide unambiguous signatures for such an origin. If future observations with high spectral resolution would confirm the predicted X-rays, the combined analysis of the X-ray and gamma-ray data will set important constraints on the origin of the accelerated particles and their interaction model. ", "introduction": "Strong gamma-ray emission in the 3-7 MeV range has been detected from the Orion molecular cloud complex with the COMPTEL instrument on the {\\it Compton Gamma Ray Observatory (CGRO}; Bloemen et al. 1994, 1997). As the observed spectrum exhibits characteristic structures (Bloemen et al. 1997), this emission is most likely due to the superposition of nuclear gamma-ray lines, mainly the 4.44 MeV line from $^{12}$C and the 6.13, 6.92 and 7.12 MeV lines from $^{16}$O. Such line emission can only be produced by accelerated particle interactions. Gamma-ray emission at photon energies $>$30 MeV was also observed from Orion, with the EGRET instrument on {\\it CGRO} (Digel, Hunter, \\& Mukherjee 1995). This gamma-ray emission is consistent with pion production and bremsstrahlung due to irradiation by standard Galactic cosmic rays (Digel et al. 1995). As such cosmic rays underproduce the observed line emission by at least three orders of magnitude, the gamma-ray line production in Orion must predominantly be a low energy cosmic ray phenomenon. Information on the spatial distribution of the gamma-ray line emission in Orion has come from both the COMPTEL and {\\it CGRO}/OSSE observations. The extended nature of the emission seen in the COMPTEL map of Orion (Bloemen et al. 1997) could provide an explanation for the fact that so far it was not possible to confirm the COMPTEL results with OSSE (Murphy et al. 1996; Harris et al. 1998). Based on the observed line widths, Bloemen et al. (1994) first suggested that the line emission is produced by accelerated C and O ions interacting with ambient H and He, rather than by accelerated protons and $\\alpha$-particles interacting with ambient C and O. More detailed analyses of the initial COMPTEL data have shown that a mix of the two processes could not be ruled out (Ramaty, Kozlovsky, \\& Lingenfelter 1995; Cowsik \\& Friedlander 1995). But, as the emission peaks in the more recent COMPTEL data do not appear at the line center energies for $^{12}$C and $^{16}$O de-excitations (Bloemen et al. 1997), a significant narrow-line contribution from accelerated proton and $\\alpha$-particle interactions seems to be excluded (Kozlovsky, Ramaty, \\& Lingenfelter 1997). This conclusion is also supported by energetic arguments, as the very large power deposited by the accelerated particles into the ambient medium in Orion is lowered by enhancing the C-to-proton and O-to proton abundance ratios (Ramaty et al. 1995; Ramaty, Kozlovsky \\& Lingenfelter 1996). Apart from the observed emission in the 3-7 MeV band, the COMPTEL observations revealed only upper limits at other gamma-ray energies (Bloemen et al. 1994, 1997). In particular, the upper limit on the 1-3 MeV emission sets constraints on the accelerated Ne-Fe abundances relative to those of C and O. The suppression of both the Ne-Fe and proton and $\\alpha$-particle abundances relative to C and O could be understood if the seed particles injected into an as-yet unknown particle accelerator (see Nath \\& Biermann 1994; Bykov \\& Bloemen 1994) come from the winds of massive stars or the ejecta of supernovae resulting from massive progenitors (Bykov \\& Bloemen 1994; Ramaty et al. 1995; Cass\\'e, Lehoucq, \\& Vangioni-Flam 1995; Ramaty et al. 1996; Parizot, Cass\\'e, \\& Vangioni-Flam 1997a). Ip (1995) and Ramaty et al. (1996) have also considered the possible acceleration of ions resulting from the breakup of interstellar dust. The gamma-ray line production in Orion should be accompanied by a large ionization rate of the ambient medium which could exceed the observed infrared luminosity (Cowsik \\& Friedlander 1995). This problem is alleviated if the gamma-rays are produced at cloud boundaries, but not in their interiors. The accelerated particles could have ionized $\\sim$2$\\times$10$^4$M$_\\odot$ in 10$^5$ years (Ramaty 1996), a small fraction of the total available mass. It is thus possible that a large fraction of the power that accompanies the gamma-ray production is deposited in an ionized gas. While the X-ray emission produced by low energy particle interactions is potentially a promising tracer of low energy cosmic rays in the Galaxy (e.g. Hayakawa \\& Matsuoka 1964), there are as-yet no astrophysical X-ray observations that unambiguously indicate the presence of such cosmic rays. The Orion region, however, has become an interesting target owing to the COMPTEL discovery of the nuclear gamma-ray line emission. A variety of processes lead to X-ray production by low energy ion interactions. Inverse bremsstrahlung (Boldt \\& Serlemitsos 1969) results from the interactions of fast ions and ambient electrons; secondary electron bremsstrahlung is produced by knock-on electrons accelerated in fast ion interactions (Hayakawa \\& Matsuoka 1964). Both of these processes lead to continuum X-ray emission. X-ray line emission results from atomic de-excitations in the fast ions following electron capture (Silk \\& Steigman 1969; Watson 1976; Pravdo \\& Boldt 1975; Bussard, Ramaty, \\& Omidvar 1978) and in ambient ions following inner-shell vacancy creation. The latter process has not yet been applied to astrophysics. Dogiel et al. (1997) have recently considered the X-ray emission that should accompany the gamma-ray line production in Orion. They have only considered the secondary electron bremsstrahlung and concluded that the 0.5-2 keV emission that accompanies the observed gamma-ray line emission from Orion will exceed the upper limits that they derived using ROSAT observations. We have subsequently taken into account both continuum processes and line emission from de-excitations in fast O (Ramaty, Kozlovsky, \\& Tatischeff 1997a) and showed that, even though the inverse bremsstrahlung is more important than the secondary electron bremsstrahlung, the total X-ray continuum emission from Orion is not inconsistent with the Dogiel et al. (1997) derived ROSAT upper limit. On the other hand, we showed that a conflict may exist between that ROSAT upper limit and the X-ray line emission following electron capture onto fast O nuclei. However, as we suggested, this conflict could be resolved if the X-ray and gamma-ray lines are produced in an ionized medium or if the current epoch accelerated particle spectrum is suppressed at low energies, for example by energy losses. In this paper we present detailed calculations of X-ray continuum and line production by accelerated particle interactions. The bulk of our treatment is for a steady state, thick target model with a neutral ambient medium. This is the standard model in which most of the gamma-ray calculations have been carried out (e.g. Ramaty et al. 1996). But we have also investigated the effects of an ionized ambient medium and a time-dependent model, as these modifications could have important consequences on the predicted X-ray to gamma-ray production ratio. In our treatment of the continuum, we have supplied the details of the calculations and we have improved the employed cross sections, thereby confirming our previous preliminary results (Ramaty et al. 1997a). We have greatly expanded our treatment of X-ray line emission. We have investigated in detail the atomic physics relevant to line emission from de-excitations in fast O, checking our theoretical calculations against laboratory data whenever available. We then expanded the treatment to the other abundant accelerated ions (C, N, Ne, Mg, Si, S and Fe), and we have also calculated the X-ray line emission produced in ambient ions following inner-shell vacancy creation by the accelerated particles. We have used the ROSAT all-sky survey (Snowden et al. 1995) to derive the X-ray count rates from the Orion region that could be associated with accelerated particle interactions; the implied fluxes are quite different from the upper limit given by Dogiel et al. (1997). The unambiguous future detection of the predicted X-rays produced by accelerated particles in Orion, and potentially elsewhere in the Galaxy, should provide important new insights into the origin of the low energy cosmic rays whose presence in Orion is revealed by the COMPTEL gamma-ray line observations. ", "conclusions": "We have investigated all the processes that lead to X-ray production by low energy cosmic rays for a variety of accelerated particle compositions and energy spectra. We demonstrated that the dominant continuum producing process is inverse bremsstrahlung produced by fast ions interacting with ambient electrons. In addition, there is also a significant contribution from the bremsstrahlung produced by secondary knock-on electrons. However, below a few keV the total X-ray emission produced by accelerated ions is dominated by relatively broad line emission (line widths $\\delta E/E$$\\simeq$0.1) resulting from de-excitations in the fast ions following electron captures and excitations. In addition, accelerated particle interactions also produce much narrower X-ray lines, due to inner-shell vacancy creation. The most prominent of such line is that at 6.4 keV from ambient Fe. We have calculated the X-ray line and continuum emission produced by the accelerated particles in Orion which are thought to be responsible for the nuclear gamma-ray line emission observed with COMPTEL (Bloemen et al. 1994, 1997). By first comparing the results with the extragalactic diffuse X-ray background, we found that while the continuum is generally below this background, the line emission from about 0.5 to 1.5 keV exceeds the background for all the combination of parameters that we considered. We wish to point out that there could be a significant contribution to the $\\sim$0.5-1.5 keV diffuse X-ray background from as-yet unknown sources within our Galaxy (e.g. Park et al. 1997), leaving the possibility that a substantial fraction of the observed X-ray intensity in this energy range results from low energy cosmic ray interactions. We have also compared our results with ROSAT observations of Orion in the 0.47 to 1.2 keV energy band, again normalizing the X-ray emission to the observed gamma-ray emission. We found that there is no conflict between the predicted total X-ray emission (lines and continuum) and the data for a broad range of parameters if the gamma-ray line emission is uniformly distributed over the entire molecular cloud complex. This conclusion differs from our previous one (Ramaty et al. 1997a) because of a lower predicted X-ray line emission, resulting from improved atomic physics input, and because our estimated ROSAT flux from Orion is higher than the upper limit given by Dogiel et al. (1997). However, the COMPTEL data show significant spatial structure. We found that for the most prominent hot spot in the COMPTEL map, the standard thick target, steady state interaction model, with a neutral ambient medium, predicts X-ray fluxes which exceed the ROSAT data for a broad range of parameters. But the calculations could be consistent with the data for any one, or a combination of the following possibilities: a very hard accelerated particle spectrum; a partially ionized ambient medium; and a time-dependent accelerated particle energy spectrum resulting from essentially instantaneous acceleration some tens of thousand of years ago. There are as-yet no astrophysical X-ray observations that would unambiguously indicate an origin resulting from low energy, accelerated ion interactions. Our calculations show that the most promising signatures are the relatively broad lines between 0.5 and 1.5 keV, mainly the lines from fast O, and that a promising target is the Orion region where the presence of such accelerated particles is known from gamma-ray line observations. We acknowledge K. Omidvar for discussions on the atomic processes leading to X-ray line production. V. T. acknowledges an NRC-NASA/GSFC Research Associateship. \\clearpage" }, "9804/astro-ph9804025_arXiv.txt": { "abstract": "We report the serendipitous discovery of a 7-s X-ray pulsar using data acquired with the {\\it Advanced Satellite for Cosmology and Astrophysics} (\\asca ). The pulsar is detected as an unresolved source located towards a region of the Galactic plane ($l,b \\simeq 29.5, 0.08$) that coincides with an overdensity of star-formation tracers. The signal suffers tremendous foreground absorption, equivalent to $N_H \\simeq 10^{23}$ cm$^{-2}$; the absorption correlates well with a line-of-sight that is tangential to the inner spiral arms and the 4-kpc molecular ring. The pulsar is not associated with any known supernova remnants or other cataloged objects in that direction. The near sinusoidal pulse (period $P \\simeq 6.9712$) is modulated at 35\\% pulsed amplitude, and the steep spectrum is characteristic of hot black-body emission with temperature $kT \\sim 0.65$ keV. We characterize the source as an anomalous X-ray pulsar (AXP). ", "introduction": "A canonical young pulsar (period $\\sim 100$ ms and stellar dipole field $\\sim 3\\times10^{12}$ G) is a rapidly rotating neutron star, created as the stellar remnant during a Type II (or Ib) supernova explosion of a massive star. The birthrate of pulsars is known to be close to $1 - 3$ per century, and it is estimated that there are about $10^5$ active and $10^8$ defunct neutron stars in the Galaxy (see Lorimer et al. 1993 and refs. therein). In the last few years, there has been growing recognition of a population of ultra-magnetized neutron stars, or ``magnetars'' (Thompson \\& Duncan 1993). The mostly circumstantial evidence comes from investigations of the following categories of objects: the soft gamma-ray repeaters (Thompson \\& Duncan 1995; Frail et al. 1997), long period pulsars in supernova remnants (Vasisht \\& Gotthelf 1997 and refs. therein), other seemingly isolated, young, long period pulsars (Thompson \\& Duncan 1996) nowadays referred to as the anomalous X-ray pulsars (AXP; van Paradijs et al. 1995), and perhaps their older variants (Kulkarni \\& van Kerkwijk 1998). These objects share some common properties; they are steady, bright X-ray sources ($L_X \\simgt 10^{35}$ erg s$^{-1}$) which show no signs for an accompanying companion, those with known periods are found to be spinning down, and all are relatively young ($ \\simlt 10^{5}$ yr-old). The evolutionary consequences of such large dipole fields are reflected in the properties listed above. Most importantly, large braking torques acting on the star cause it to spin-down rapidly, and the magnetic free energy quickly dominates over the rotation energy, i.e., within several hundred years. For the above sources, the rotation rates lie between 6 - 12 s, with ages $\\simlt 10^5$ yr (for the SGRs the evidence for periods is indirect, however, their ages are well constrained due to their association with supernova remnants). It is believed that field decay, which is expected for ultramagnetized neutron stars (Thompson \\& Duncan 1996; also Goldreich \\& Reisenegger 1992), influences the thermal evolution and powers the large X-ray luminosities observed for the purported magnetars, $L_X \\simgt 10^{35}$ erg s$^{-1}$. If magnetars represent the tail-end of the magnetic field distribution of neutron stars, then they are bound to be rare. Assume that their birthrate is 10\\% the birthrate of neutron stars (some justification for this comes from the estimated birthrates of SGRs; Kulkarni et al. 1994), and that they have active X-ray lifetimes of $\\sim 10^{5}$ yr. These assumptions imply that at present there are only $\\sim 100$ active magnetars in the Galaxy, a conclusion that is borne out by the observations of the aforementioned objects. The fact that we observe the five known AXPs through large column densities in the Galaxy, $N_H \\simgt 10^{22}$ cm$^{-2}$, suggests that they are indeed that rare, and the fact that they are often associated with supernova remnants or lie near star-formation regions (in spite of the large random velocities usually attributed to neutron stars; Lyne \\& Lorimer 1994) suggests that they are young. Similarly, only two Galactic SGRs are known, and it has been suggested that the SGR population census is nearly complete (Kouveliotou 1995). In summary, AXPs have long rotation periods, hot blackbody-like spectra ($kT \\sim$ 0.5 keV) with $10^{35-36}$ erg s$^{-1}$ steady luminosities, and have thus far only been observed at X-ray wavelengths. A search for new AXPs in the ASCA database has turned up another candidate, which we refer to as \\psr. ", "conclusions": "On the basis of its long rotation period, steady X-ray flux, steep spectral characteristics, and location in the Galactic plane ($|b| \\le 0.5$), we classify \\psr\\ as an anomalous X-ray pulsar (see Table 1). The high foreground absorption suggests that the pulsar is distant, and its line of sight along the tangent to the Sagittarius-Carina and Scutum-Crux spiral arms, and the 4-kpc molecular ring justifies its enormous foreground absorption (see figure 4). Its $N_H$ is roughly twice that of the nearby remnant Kes 73 for which quoted distances lie between $10 - 20$ kpc (Blanton \\& Helfand 1996). However, the disparity in $N_H$ does not in itself imply a great dissimilarity in distances. For instance, at 10 kpc the lines of sight vectors to these two objects are already separated by $\\sim 100$ pc, the typical sizes and scale heights of dense giant molecular clouds which are likely to be responsible for most of the absorbing gas. For the purposes of this article we assume the distance to be 15 kpc, an estimate likely to be accurate to within a factor of two. The steep X-ray spectrum is characteristic of the Wien tail of a blackbody radiator. Using the best fit blackbody parameters the isotropic X-ray luminosity is $L_X \\simeq 2.5\\times 10^{35}d_{15}^2$ erg s$^{-1}$, the distance being 15$d_{15}$ kpc. In effect, the X-ray pulsations can be ascribed to the viewing of a rotating stellar hotspot of area $0.15 A_sd_{15}^2$, where $A_s$ is the area of a neutron star of radius 10 km; this estimate ignores any relativistic corrections to the inferred area. Note that the spectrum is unlike that of any accreting high-mass neutron star binary. Although such binaries have periods in the range 0.07 - 900 s, and sometimes go into low luminosity states with $L_X \\sim 10^{35}$ erg s$^{-1}$, they generally display very hard spectra ($0.8 < \\Gamma < 1.5$), and show stochastic variability on all time-scales (Nagase 1989; Koyama et al. 1989) as is generally seen in accretion powered sources. We find no evidence for such variability in our data. With an AXP classification in hand we can compare the properties of \\psr\\ with those of five other members of the AXP family in Table~1. A few years ago Schwentker (1994) reported weak 5-s pulsations from RX J1838.4$-$0301 which have not yet been confirmed, Mereghetti et al. (1997) have argued that this X-ray source might be due to coronal emission from a late type star. We, therefore, exclude this source from our list. Although only a future $\\dot P$ measurement can help determine the linear spindown age of \\psr\\ (an estimator for the age of an isolated neutron star), its location in the Galactic plane suggests that it is young, $\\tau < 10^5$ yr. We consider it extremely likely that the pulsar is associated with one of the several star-formation complexes expected to lie along this line-of-sight (see Fig 4), out to a distance of 20 kpc. The pulsar lies along a rich region of the Galaxy; there are 10 supernova remnants, several radio pulsars, and the $\\gamma$-ray source GRO J1838$-$04, all within a $3\\times3$ deg$^2$ patch of sky surrounding \\psr. However, to the best of our knowledge no cataloged sources are associated with it. We now describe other models that address the unique properties of AXPs. It is often noted in the literature, that the inferred accretion rate for the pulsars in Table~1 are close to those expected for accretion powered pulsars with field strengths $B \\sim 10^{11-12}$~G, spinning at their equilibrium periods $P_{eq}$ (see Bhattacharya \\& van den Heuvel 1991). This motivated Mereghetti \\& Stella (1995) to suggest that these pulsars are members of a subclass of low mass X-ray binaries (LMXBs) in equilibrium rotation, with the stellar magnetic field of order $B_s \\sim 10^{11}$ G. In contrast, van Paradijs et al. (1995) argue that these objects are isolated neutron stars accreting from a fossil disk, while Ghosh et al. (1997) suggest that AXPs are formed as the result of a Thorne-$\\dot{\\rm{Z}}$ytkov phase of a high mass X-ray binary with strong spherical accretion leading to the soft X-ray spectra with high foreground absorption. It is worth mentioning that accretion scenarios would be hard-pressed to explain the spin-down age of at least one member of Table 1, the $\\sim 2000$ yr of the pulsar in Kes~73 (see Table 1; Gotthelf \\& Vasisht 1997). First, it is difficult for accretion torques to spin-down a pulsar to 12-s in $\\sim 10^3$ yr from initial periods $P_i \\simlt 10^2$ ms unless, of course, the pulsar were born a very slow rotator, which is quite interesting in its own right. Secondly, if the pulsar were rotating near its equilibrium period, as in the Ghosh and Lamb (1979) scenario, the spin-down time of ${P/ 2\\dot P} \\sim 3900$ yr is inconsistent with the luminosity implied accretion rate, $\\dot M \\simeq 10^{-11}$ M$_\\odot$ yr$^{-1}$ (assuming the pulsar has a standard dipolar field $\\simeq 10^{12}$ G); these usually lie in range $10^4 - 10^5$ yr. In conclusion, further X-ray observations are required to secure the classification of the 7-s pulsar to the growing family of AXPs - by measuring the long term stability of the X-ray flux, secular trends in the pulse period including Doppler modulation, the lack of which will firm up the likelihood against an accreting binary hypothesis. Indeed, if AXPs are akin to SGRs then they could display sporadic hard X-ray transients, although such behavior is yet to be observed. Infrared observations to search for a possible counterpart to \\psr\\ could be carried out in spite the somewhat crude localization; we mention that past optical/IR searches for counterparts have been unsuccessful. Furthermore, radio observations to uncover an associated supernova remnant could be pursued. The high foreground absorption could easily cloak the soft X-ray emission of an aged, few$\\times 10^4$ yr-old, remnant. We find it remarkable that the period of \\psr\\ agrees so well with that of 1E~2259+586, although at present we believe this is sheer coincidence." }, "9804/astro-ph9804213_arXiv.txt": { "abstract": "In efforts to demonstrate the linear Hubble law $ v = H r $ from galaxy observations, the underlying simplicity is often obscured by complexities arising from magnitude-limited data. In this paper we point out a simple but previously unremarked fact: that the shapes and orientations of structures in redshift space contain in themselves independent information about the cosmological redshift-distance relation. The orientations of voids in the CfA slice support the Hubble law, giving a redshift-distance power index $ p = 0.83 \\pm 0.36 $ (void data from Slezak, de Lapparent, \\& Bijoui 1993) or $ p = 0.99 \\pm 0.38 $ (void data from Malik \\& Subramanian 1997). ", "introduction": "Hubble's (1929) observation that redshift increases linearly with distance for nearby galaxies has now been known for almost seventy years, and it is likely that its validity is not doubted by many. Yet, any attempt to demonstrate this simple law from galaxy observations soon descends into the complexities of magnitude-limited observations, the broad galaxy luminosity function, Malmquist bias, and, at larger redshifts, $K$-corrections, evolution, etc. This has permitted an often-ignored but persistent challenge to a linear redshift law from those preferring a quadratic relation, with both challengers and supporters citing data ranging from relatively nearby, bright optical galaxies (Soneira 1979, Segal 1980) through the {\\it IRAS} 1.2~Jy redshift catalog (Segal et al. 1993; Koranyi \\& Strauss 1997). One complication in understanding the classical Hubble diagram has been inhomogeneity, appearing as a dependence of density on radius or redshift or in clustering and peculiar velocities. With the advent of more galaxy redshift catalogs covering larger amounts of the sky, we obtain ever clearer pictures of the universe. The CfA ``slice'' (de Lapparent, Geller, \\& Huchra 1986), in particular, first revealed dramatic structures, large voids separated by well-defined walls, extending to a significant fraction of the survey volume, perhaps even calling into question whether a survey to this depth represents a statistically fair sample of the universe. In other ways, however, this inhomogeneity itself can be useful. In this paper we use the shapes and orientations of structures to obtain information about the background cosmology in which they are embedded. In Section 2 below we discuss how shapes and orientations of structures in redshift space depend on the redshift-distance relation, and in Section 3 we apply these considerations to the CfA slice (de Lapparent, Geller, \\& Huchra 1986) and the Las Campanas redshift survey (Shectman et al 1996, LCRS). Section 4 contains a final discussion. ", "conclusions": "We have shown in this paper how the apparent shapes and orientations of objects in redshift space can be used to determine the redshift-distance relation. When redshift is not linear in distance, objects such as voids that are intrinsically round in space when viewed in redshift space are stretched along the line of sight; and an initially isotropic distribution of orientations becomes distorted in the radial direction, with measurable effect, as in \\eq{mu2d}. Application to voids in the CfA slice gives redshift power index $ p = 0.83 \\pm 0.36 $ from Slezak et al. (1993) and $ p = 0.99 \\pm 0.38 $ from Malik \\& Subramanian (1997), both a modest preference for $ p = 1 $ over $ p = 2 $. To obtain these results in redshift space with a true $ p = 2 $, voids in the CfA slice in space would have to be flattened and preferentially aligned transverse to the line of sight. The main complication to the simple interpretation is likely to come from peculiar velocities; an expansion velocity will introduce an apparent line-of-sight elongation. However, this effect is expected to be small: in the most extreme case, of an uncompensated, completely empty, isolated void in an $ \\Omega_0 = 1 $ universe, expansion would make $ p = 1 $ appear to be $ p = \\case43 $. In any case, unless voids are contracting, a condition difficult to make physical sense of, peculiar velocities will only increase, not decrease, the apparent value of $p$, and will not distort a quadratic redshift law to appear to prefer $ p = 1 $. We do not expect void orientation statistics to replace classical methods for demonstrating the Hubble law. The most precise confirmation of the linearity of the Hubble law, $ p/5 = 0.2010 \\pm 0.0035 $, has been obtained recently from the Hubble diagram of type Ia supernovae (Riess, Press, \\& Kirshner 1996). The luminosity function as a function of redshift and the radial dependence of galaxy density in the {\\it IRAS} 1.2-Jy catalog (Koranyi \\& Strauss 1997) also support the linear Hubble law. However, our method offers an independent verification of the Hubble law in which magnitude-limited data, Malmquist bias, and evolution do not present problems. Further examination of the effect in real data, especially including in detail the peculiar velocities of the void walls, will undoubtedly introduce complications to the simple relations in equations (\\ref{mu3d}) and (\\ref{mu2d}). Whatever these complications, they will be different from those that enter arguments over the redshift-magnitude relation." }, "9804/astro-ph9804163_arXiv.txt": { "abstract": "We present radio observations of the gamma-ray burster \\grb\\ made with the Very Large Array (VLA) and the Owens Valley Radio Observatory (OVRO) spanning a range of postburst timescales from one to 300 days. A search for a time-variable radio source was conducted covering an area which included a fading X-ray source and an optical transient, both of which are thought to be the long wavelength counterparts to the gamma-ray burst. At the position of the optical transient sensitive limits between 10 $\\mu$Jy and 1 mJy can be placed on the absence of a radio counterpart to \\grb\\ between 1.4 and 240 GHz. We apply a simple formulation of a fireball model which has been used with some success to reproduce the behavior of the optical and X-ray light curves. Using this model we conclude that the radio non-detections are consistent with the peak flux density of the afterglow lying between 20-40 $\\mu$Jy and it requires that the optical flux peaked between 4 and 16 hours after the burst. ", "introduction": "The gamma-ray burst of 28 February 1997 was a turning point in our understanding of these enigmatic objects, resulting in the first-ever discovery of X-ray and optical counterparts to a burst. Within the original 3-arcminute localization provided by the Wide Field Cameras (WFC) on board the BeppoSAX satellite, a previously unknown X-ray source 1SAX\\ts{J0501.7+1146} was detected by the Narrow Field Instruments (NFI). Eight hours after the burst the flux of 1SAX\\ts{J0501.7+1146} was 2.8$\\times{10}^{-12}$ erg cm$^{-2}$ s$^{-1}$ (2-10 keV), but three days later its flux had dropped by a factor of 20 (Costa et al. 1997). A comparison of V- and I-band images taken on 28 February and 8 March revealed an optical transient within a reduced error box, defined by the intersection of the WFC circle, the $\\pm$50\\arcsec\\ circle of the NFI, and the Interplanetary Network (IPN) annulus (van Paradijs et al. 1997, Hurley et al. 1997). Predictions of long-lived afterglows from gamma-ray bursts at X-ray, optical and radio wavelengths have been made for some time (e.g. Paczy\\'nski \\& Rhoads 1993, M\\'esz\\'aros, Rees, \\& Papathanassiou 1994, Katz 1994, M\\'esz\\'aros \\& Rees 1997). In particular, a power-law decay in the observed long-wavelength flux with time is a generic consequence of a class of models known as ``fireballs''. The gamma-rays are thought to be produced when the relativistically expanding blast wave (aka fireball) is slowed down by the ambient gas or is self-shocked by its own ejecta. The fireball accelerates particles in a shock which then radiate via the synchrotron process (M\\'esz\\'aros et al. 1994, Waxman 1997a, 1997b, Sari, Piran, \\& Narayan 1998). As the fireball expands it cools, shifting the peak in the spectrum to lower energies and resulting in delayed emission at longer wavelengths. The optical and X-ray decay from \\grb, as well as that from several other subsequent GRBs, is consistent at least to first order with one of the simpler formulations of these models (M\\'esz\\'aros \\& Rees 1997). Costa et al. (1997) fit a t$^\\delta$ decay to the X-ray data with $\\delta\\simeq-1.33\\pm{0.12}$, while global fits to the X-ray, optical and infrared data give $\\delta=-1.2$ (Wijers, Rees and M\\'esz\\'aros 1998) and $\\delta=-1.09\\pm0.23$ (Reichart 1997). More recent optical fits, aided by a long time-baseline, yield $\\delta=-1.12\\pm0.08$ (Garcia et al. 1997), $\\delta=-1.21\\pm0.02$ (Masetti et al. (1997), and $\\delta=-1.10\\pm0.04$ (Galama et al. 1998). The character of this decay agrees well with a fireball produced by a one-time impulsive injection of energy in which only the forward blast wave efficiently accelerates particles (i.e. the adiabatic piston model). The adiabatic model predicts a simple relation between the slope of the temporal decay $\\delta$ and the flux spectrum $\\beta$=${2\\over3}\\delta$. Observationally the spectrum is not well determined, with $\\beta=-0.7\\pm0.1$ optically (van Paradijs et al. 1997) and X-ray values of $\\beta\\simeq-0.9$ with large scatter (Frontera et al. 1998). The slope of the particle spectrum $p$ (where $\\beta=(p-1)/2)$ inferred from these values of temporal decay is $p\\sim-2.6$, not an unreasonable value for a relativistic shock (Blandford and Eichler 1987). Given the early success of the fireball model in predicting the gross properties of the optical and X-ray behavior of \\grb, it is reasonable to look for delayed radio emission from this burst. This is particularly relevant in the light of the discovery of the radio afterglow from GRB\\thinspace{970508} (Frail et al. 1997a) which exhibited temporal and spectral behavior consistent with this model (Waxman, Kulkarni \\& Frail 1997). The properties of the fireball for \\grb\\ are well constrained by the optical and X-ray data. Thus the presence or absence of radio emission at late times has a bearing on the validity of this model. With this in mind, we began a radio search centered on the optical transient detected by van Paradijs et al. (1997). This {\\it Letter} is a summary of our monitoring program for the first 300 days. ", "conclusions": "The detection of an optical transient, thought to be the afterglow from \\grb, has made it possible to perform a search for time-variable radio emission at the optical position. Delayed emission at longer wavelengths (X-ray, optical and radio) is a generic prediction of all fireball models. VLA and OVRO observations have been made that span a range of postburst timescales from one to 300 days putting limits on the absence of a radio counterpart to \\grb\\ between 10 $\\mu$Jy and 1 mJy. Applying a simple version of the fireball model which has been used successfully to fit the temporal behavior of the decaying X-ray and optical emission from \\grb\\ suggests that the radio afterglow has yet to peak. If this is correct then continued deep imaging in the coming months offers the possibility for the detection of a weak but increasing radio signal. Detecting this emission would be a powerful verification of the fireball model as described by M\\'esz\\'aros \\& Rees (1997). Indeed, as the recent detection of GRB\\thinspace{970508} has taught us, radio afterglows yield unique GRB diagnostics that are not obtainable by any other means (Frail et al. 1997b). Unlike optical or X-ray wavelengths one is presented with the possibility of following the {\\it full} evolution of the fireball emission through all its different stages; first while it is optically thick, then as it slowly rises to a peak flux density and thereafter decays, making a transition from an ultra-relativistic to sub-relativistic shock. Furthermore, both the scintillation of the radio source (Goodman 1997) and its flux density, when it is synchrotron self-absorbed (Katz 1994), allow a determination of the size and expansion of the fireball. If the radio afterglow is going to be detected from \\grb\\ and others like it, then continued long-term monitoring is going to be required at the microJansky level." }, "9804/astro-ph9804040_arXiv.txt": { "abstract": " ", "introduction": "Few properties of astronomical masers are determined directly by observations, most are inferred indirectly. Foremost among the latter is the maser saturation stage. Saturation has a significant impact on maser growth, so determining whether a maser is saturated ($J > J_s$, where $J$ is the angle-averaged intensity and $J_s$ is the saturation intensity) or not is usually a precondition for analysis of the observations. Unfortunately, this crucial issue is not convincingly settled. Strong masers are generally believed to be saturated, but in most cases the evidence is less than compelling as it relies primarily on plausibility arguments rather than quantitative tests (see e.g.\\ \\[Book], \\S 8.6). This unsatisfactory situation reflects a fundamental difficulty --- neither $J$ nor $J_s$ is directly measurable. The saturation parameter $J_s$ is a theoretical quantity, determined only within the context of a given pumping scheme. And because maser radiation is highly beamed, $J = I\\Omega/4\\pi$ so this quantity, too, cannot be directly measured; the intensity $I$ is measurable when the maser is resolved, but the beaming angle $\\Omega$ is unobservable. Similarly, the amplification optical depth has never been directly determined for any maser that amplifies its own radiation. Recent VLA observations of OH 1720 MHz masers near the Galactic center by \\[YZ96] open up new possibilities for direct determination of some maser properties. Significant circular polarization (upward of 20\\%) is detected in various spectral features, and the right- and left-hand components coincide on the sky, as expected from the Zeeman effect. Furthermore, the spectral shape of the Stokes parameter $V$ follows an antisymmetric S-curve with sharp reversal at line center, the typical profile for Zeeman shift \\DnuB\\ much smaller than the Doppler linewidth \\DnuD. Similar results were previously reported for H$_2$O masers in star-forming regions by \\[FbG] and for OH 1612 MHz masers in OH/IR stars by \\[ZFix], but the polarization was lower and the quality of the data not nearly as high. The general maser polarization solution was recently derived for arbitrary values of \\eq{ \\xb = {\\DnuB \\over \\DnuD} } (Elitzur 1996, hereafter \\[E96]) and the solution properties at $\\xb \\ll 1$ closely match the observed circular polarization. Here I show that a comparative analysis of the spectral profiles of $I$ and $V$, two measurable independent maser intensities, offers direct determination of various maser properties, in particular the saturation stage. The analysis is readily performed with the aid of the ratio profile \\eq{\\label{R0} \\R(\\nu) = {V(\\nu) \\over I'(\\nu)} = {v(\\nu)\\over I'(\\nu)/I(\\nu)} } where the prime denotes derivative with respect to frequency and $v = V/I$ is the fractional circular polarization. When $\\xb \\ll 1$, spectral analysis of \\R\\ offers intrinsic sensitivity of order \\xb\\ and has long been an important tool in studies of the Zeeman effect of thermal radiation (see e.g.\\ \\[Trol]). In that case \\R\\ is constant across the spectral line and its magnitude determines the magnetic field along the line of sight. This constancy of \\R\\ follows from some simple, general symmetry arguments as shown by \\[Cru] (see also \\S3 below). However, maser exponential amplification during unsaturated growth destroys both the underlying symmetry and the constancy of \\R, the saturation process restores both. The key to the different behavior, and \\R-profiles, in the two regimes is the narrowing of the maser line during unsaturated amplification and its rebroadening during saturation. The important differences between thermal and maser polarization are discussed in detail below. For completeness, some basic elements of the polarization theory developed in \\[E96] are reproduced in \\S2. The \\R\\ profile is discussed in \\S3 for thermal radiation and in \\S4 for maser radiation when $\\xb \\ll 1$. In \\S5, circular polarization for fully resolved Zeeman patterns, $\\xb > 1$, is discussed. The implications for observations are discussed in detail in \\S6. ", "conclusions": "" }, "9804/astro-ph9804089_arXiv.txt": { "abstract": "Stellar dynamics is almost unreasonably well suited for an implementation in terms of special-purpose hardware. Unlike the case of molecular dynamics, stellar dynamics deals exclusively with a long-range force, gravity, which leads to a computational cost scaling as the square of the number of stars involved. While special tricks can lead to a reduction of this cost from $\\sim N^2$ to $\\sim N\\log N$ in the case of very large particle numbers, such tricks are not suitable for all areas within stellar dynamics. When a stellar system is close to equilibrium, and has a very high density, it still pays to compute all interactions on a star by star basis, even for $N=10^5$. Any $cN\\log N$ approach would either gloss over the subtle net effects of near-canceling interactions, driving the evolution of such a system, or would carry a prohibitively large coefficient $c$. This paper presents a brief introduction to the stellar dynamics of dense stellar systems, aimed at researchers using special purpose computers in other branches of physics. ", "introduction": "Stellar dynamics is the branch of astrophysics that studies the structure and evolution of collections of stars, from small groups to larger star clusters to entire galaxies and clusters of galaxies. The interactions between the individual stars can be modeled to a high degree of accuracy as Newtonian gravitational interactions between point masses. Only under extremely high densities, such as occurs in the nuclei of galaxies and the centers of the densest star clusters, do stars have a reasonable chance to undergo a physical collision during their life time. In contrast, a typical star, such as our own Sun, has a probability of only 1 in $10^{8}$ to undergo a collision with a neighboring star, during the remaining $5\\times10^9$ years of its life. The most spectacular example of a dense stellar system within our own galaxy is the agglomeration of roughly a million stars within the inner parsec from the center (a parsec is a unit of length, equal to a few light years, and corresponds to a typical distance between stars in the solar neighborhood). These stars describe orbits around the black hole that resides in the very center of our galaxy. The black hole itself has a mass that is a few million times larger than the mass of the Sun. The density of stars around the black hole is a million times larger than the stellar density in a typical part of the galaxy, such as where we reside. A detailed stellar dynamical modeling of the center of our galaxy is still difficult, partly because the observations of this heavily obscured area have only recently become accurate enough to tell us the physical characteristics of the system, partly because of the interference of other physical effects, such as the presence of gas clouds and ongoing star formation. Before tackling the stellar dynamics of the nucleus of our galaxy, it is therefore prudent to start our attempts with a simpler system, such as is provided by the core of a dense globular cluster. While most of the stars in and around our galaxy are spread out throughout the disk, and to a lesser extent through the halo, there are more than one hundred isolated star clusters circling the galaxy, each containing of order $10^6$ stars. In a dozen or so of those globular clusters, as they are called because of their appearance, the central densities rival that of the density in the nucleus of our galaxy. However, the absence of a large black hole, as well as gas clouds and concomitant star formation, makes it far easier to study and model globular cluster cores in detail. In addition, recent observations, notably with the Hubble Space Telescope ({\\it cf.} \\cite{Guh96}), have resolved those cores into individual stars, something that has not been possible with ground-based observations. This paper sketches some of the progress made in the study of globular cluster cores, emphasizing the role of special-purpose computers. ", "conclusions": "" }, "9804/astro-ph9804276_arXiv.txt": { "abstract": "Modelling gravity is a fundamental problem that must be tackled in $N$-body simulations of stellar systems, and satisfactory solutions require a deep understanding of the dynamical effects of softening. In a previous paper (Romeo 1997), we have devised a method for exploring such effects, and we have focused on two applications that reveal the dynamical differences between the most representative types of softened gravity. In the present paper we show that our method can be applied in another, more fruitful, way: for developing new ideas about softening. Indeed, it opens a {\\it direct\\/} route to the discovery of optimal types of softened gravity for given dynamical requirements, and thus to the accomplishment of a physically consistent modelling of disc galaxies, even in the presence of a cold interstellar gaseous component and in situations that demand anisotropic resolution. ", "introduction": "$N$-body simulations of disc galaxies rely on the use of softening. This artifice removes the short-range singularity of the gravitational interaction, which is dynamically unimportant and computationally troublesome, whereas it leaves the long-range behaviour of gravity unchanged. But softening is also a critical factor in simulations. It controls their quality and can affect their result on scales much larger than the softening length. Its dynamical effects are further exacerbated in the presence of a cold interstellar gaseous component and in situations that demand anisotropic resolution. Thus softening poses a dynamical problem of special concern, which should be probed carefully and in detail (e.g., Hernquist \\& Barnes 1990; Pfenniger \\& Friedli 1993; Romeo 1994, hereafter Paper I; Romeo 1997, hereafter Paper II% \\footnote{Sections and equations of that paper are denoted by the prefix II.}; and references therein). In Paper I, we have investigated how faithful simulations are. In particular, we have concluded that the standard way of introducing softening in the presence of stars and cold interstellar gas is definitely unsatisfactory in several regimes of astrophysical interest. It is so because important small-scale instabilities of the gaseous component, e.g.\\ those peculiar to star-formation processes, are suppressed just as unphysical noise of the stellar component. Faithfulness requires an appropriate introduction of two softening lengths, one for each component, and also a rigorous specification of the star-gas gravitational interaction. In Paper II, we have devised a method for exploring the dynamical effects of softening. As a major result, we have shown how to choose the softening length for optimizing the faithfulness of simulations to the Newtonian dynamics. Then we have focused on two applications that reveal the dynamical differences between the most representative types of softened gravity. In particular, we have concluded that it is desirable to improve the current way of introducing anisotropic softening. We need a clearer decoupling of the resolution parallel and perpendicular to the plane, and also more natural planar and vertical softening lengths. In the present paper, which completes our planned research work about softening, we propose an {\\it innovative\\/} solution to the problem. The understanding of galactic and extragalactic astrophysics is at a crucial stage. Unsolved problems are viewed in new perspectives, which promise major revisions of knowledge (see, e.g., Blitz \\& Teuben 1996; Block \\& Greenberg 1996). Recent investigations suggest, for instance, a more enigmatic interplay between stellar disc and bulge/halo (e.g., Lequeux et al.\\ 1995), a clearer relation between cold gas and dark matter in spiral galaxies (e.g., Pfenniger et al.\\ 1994; Pfenniger \\& Combes 1994; Combes \\& Pfenniger 1997), and a closer connection between the fractal structures of the interstellar medium and of the universe (e.g., de Vega et al.\\ 1996, 1998). The implications are clear: modelling gravity in $N$-body simulations of disc galaxies should offer a flexible interface with such a progress. Our solution is to optimize the fidelity of simulations to given dynamical requirements. How do we apply this idea in practice? \\begin{enumerate} \\item We impose the requirements in the wavenumber space since this is the natural dynamical domain of gravity, as Pfenniger \\& Friedli (1993) have previously emphasized. \\item We identify the softening length with the characteristic dynamical scale length. \\item Then we invert part of the method of Paper II, and the result is the optimal type of softened gravity that satisfies those dynamical requirements. \\end{enumerate} Our application covers both 2-D and 3-D modelling. The basic cases are extended to more complex situations through recipes for implementing star-gas and anisotropic softening, which have already been motivated (cf.\\ discussions of Papers I and II). Last but not least, each description is complemented by an example that leaves room for creativity. The present paper is organized as follows. The application is shown in Sects.\\ 2 and 3 (see also Appendix A), and proceeds as in the previous discussion. Comments on related works concerning softening are made in Sect.\\ 4. The conclusions and perspectives are drawn in Sect.\\ 5, where we present our three papers about softening in a more unified view and emphasize their potentially strong impact on galactic dynamics. \\begin{figure*} \\vbox{\\vspace{.1cm} \\hbox{\\hspace{-.25cm} \\psfig{figure=romeof1.ps,width=18.9cm,height=13.275cm,angle=-90}} \\vspace{-2.9cm}} \\hfill\\parbox[b]{5.7cm}{\\caption[]{Examples of 2-D modelling: {\\bf a} one-com\\-pon\\-ent case (cf.\\ Sect.\\ 2.1), {\\bf b} two-com\\-pon\\-ent case (cf.\\ Sect.\\ 2.2). The abbreviations N, T and P mean Newtonian gravity, thickness and Plummer softening, respectively}} \\end{figure*} \\begin{figure*} \\vbox{\\vspace{.1cm} \\hbox{\\hspace{-.25cm} \\psfig{figure=romeof2.ps,width=18.9cm,height=13.275cm,angle=-90}} \\vspace{-2.9cm}} \\hfill\\parbox[b]{5.7cm}{\\caption[]{Examples of 3-D modelling: {\\bf a} iso\\-tropic case (cf.\\ Sect.\\ 3.1), {\\bf b} an\\-iso\\-tropic case (cf.\\ Sect.\\ 3.2). The abbreviations N, T and P mean Newtonian gravity, thickness and Plummer softening, respectively}} \\end{figure*} ", "conclusions": "The importance of computer simulations in astrophysics is analogous to that of experiments in other branches of physics. They also serve as a welcome bridge between theories, often restricted to idealized situations, and observations, revealing instead the complexity of nature. Major present objectives are to construct physically consistent $N$-body models of disc galaxies and to simulate their dynamical evolution, especially in regimes of spiral structure in which a fruitful comparison between theories and simulations can be made (e.g., Pfenniger \\& Friedli 1991; Junqueira \\& Combes 1996; Zhang 1996; Bottema \\& Gerritsen 1997; Fuchs \\& von Linden 1998; von Linden et al.\\ 1998; Zhang 1998a, b). The construction of such models is indeed a difficult task which has not yet been fully accomplished, and which should eventually provide clues of vital importance to a number of open questions posed by both theories and observations. Our involvement has been threefold. In Paper I, we have recognized a fundamental problem posed by this research programme (for a concrete use of that analysis and for interesting remarks see, e.g., Junqueira \\& Combes 1996). In Paper II, we have devised a method for solving this problem. In the present paper, we apply this method and solve the problem, thus laying the foundations of such a plan. The {\\it major result\\/} is that gravity can be modelled so as to optimize the fidelity of simulations, and the procedure is practicable. The following conclusions point up the whys and wherefores: \\begin{enumerate} \\item Optimization is performed with respect to arbitrary dynamical requirements and, in specific examples, with respect to the Newtonian dynamics. This enriches the modelling with an {\\it unprecedented\\/} degree of freedom, which has clear epistemological motivations (cf.\\ Sect.\\ 1, discussion of the present paper). \\item Optimization is performed in the wavenumber space. This is the {\\it appropriate\\/} domain for imposing dynamical requirements on the modelling. \\item Optimization concerns {\\it both\\/} the softening length {\\it and\\/} the type of softened gravity. \\item Softening is conceived as a {\\it double\\/} artifice. The softened gravity and finite-sized particle conceptions are equivalent in the basic cases. Concerning more complex situations, the latter is particularly useful for implementing star-gas softening, whereas the former is particularly useful for implementing anisotropic softening. Thus both conceptions contribute towards the accomplishment of a physically consistent modelling. \\end{enumerate} Our application is ready for a concrete use. An attractive idea is to employ a particle-particle code together with MD-GRAPE, a highly parallelized special-purpose computer for many-body simulations with an arbitrary central force (Fukushige et al.\\ 1996). We can also employ a classical particle-mesh code. Then the dynamical effects of the grid are known and factorize as those of softening (e.g., Bouchet et al.\\ 1985; Efstathiou et al.\\ 1985; for a review see, e.g., Hockney \\& Eastwood 1988). So essentially the application proceeds as in the present paper, but it may be useful to act directly on the wavenumber space (e.g., Tormen \\& Bertschinger 1996). A more complex problem concerns tree codes, which have hierarchical structure and adaptive resolution over multiple scales (e.g., Hernquist 1987; for a review see, e.g., Pfalzner \\& Gibbon 1996). The solution to that problem would need a more advanced analysis (cf.\\ following discussion). Welcome suggestions about the choice of the code can come from cosmological simulations (e.g., Splinter et al.\\ 1998). Finally, what about the future? Our approach is connected with the technique of filtering in spectral domain used in the context of digital image processing. This is a rapidly evolving field with growing applications in science and engineering, which can promote further substantial advances in $N$-body modelling of disc galaxies. For instance, wavelets are ideal for resolving multi-scale problems in space and/or time, such as those concerning turbulence, bifurcations, fractals and many others (see, e.g., Kaiser 1994; Holschneider 1995; Bowman \\& Newell 1998; for an alternative analysis tool see, e.g., Stutzki et al.\\ 1998). Speculating further, wavelets might be used for speeding up simulations through fast solution of linear systems (cf.\\ Press et al.\\ 1992, pp.\\ 597--599 and 782). These are the merits of our contribution. We hope that the trilogy (Papers I--III) and further reflections (Romeo 1998) will encourage $N$-body experimenters to model gravity so as to optimize the fidelity of their simulations, and that the result will be a stronger interdisciplinary connection with theories and observations." }, "9804/astro-ph9804106_arXiv.txt": { "abstract": "We propose that the majority of quasars at redshift $z\\sim 1 - 5$ formed in the environment of new born collapsed halos with 1-D velocity dispersion $\\sigma_v^{1d} \\sim 400 \\kms$. The harboring coefficient $f$ of quasars per halo and the lifetime of quasars depend only on local process, not modulated by the density inhomogeneities on scales larger than the size of the halos. Thus, the bias of quasars on scale larger than the size of these halos is mainly determined by the parameter $\\sigma_v$ used for quasar environment identification. With this model, the popular structure formation models, like SCDM and LCDM, can be fairly well reconciled with the data of quasars, including a. observed feature of the environment for quasars; b. redshift evolution of quasar abundance; c. the two-point correlation functions of quasars. This bias model predicts that the correlation function of quasars doesn't significantly evolve, or only slightly increases with redshift. ", "introduction": "\\bigskip Mass distribution at high redshifts is being a hot subject of the large scale structure study. Data of various objects at moderate and high redshifts, in particular, clusters of galaxies, are becoming available for probing the formation and evolution of structures and for discriminating among popular dark matter models (e.g. Jing \\& Fang 1994; Eke, Cole \\& Frenk 1996; Bahcall, Fan \\& Cen 1997; Kitayama \\& Suto 1997). Considering that quasars are the most distant among various luminous objects, they have also been applied in this approach (e.g. Bi \\& Fang 1997). However, as a mass tracer of the cosmic matter field, quasars are still playing a role different from clusters of galaxies. The problem stems from so-called ``bias''. Clusters are a biased tracer of the mass distribution. The correlation amplitude of clusters is believed to be much higher than that of the underlying matter and increases strongly with the cluster richness (Bahcall \\& Soneira 1983). This bias is plausibly explained by the mechanism that the observed clusters are identified as massive collapsed halos of the density field (Kaiser 1984). That is, the bias of clusters is completely determined by the gravitational parameters, like mass $M$ and virial radius $r_{vir}$ used to identify the halos. With this approach, a detailed confrontation can be made between theories and the observations of clusters. Quasars may also be biased tracer of the mass distribution. Recent observations indicate that the correlation amplitude of quasars may also be different from the underlying dark matter (Mo \\& Fang 1993; Komberg, Kravtsov \\& Lukash 1994; Croom \\& Shanks 1996; Franca, Andreani \\& Cristiani, 1997). However, so far no detailed model is available for the bias of quasars, though their high clustering strength and environment imply that quasars are hosted by massive halos (see below). Because of the lack of such a model, one cannot confront the data of quasars with theoretical models in the way as for clusters. For instance, the abundance of quasars can only be used as an upper or lower limit to certain massive halos; no detailed comparison between the number densities of quasars and of halos is allowed. Obviously, it is very important to understand what kind of mass halos are associated with the majority of quasars. Such a knowledge will not only enable the observational data of quasars to be powerful tests for theoretical models of galaxy formation but also tell that what type of local environments is responsible for intriguing the nuclear activities of quasars. Like clusters and groups of galaxies, it is generally believed that quasars should be associated with certain type of collapsed dark matter halos. Yet, different from identification of clusters, the environment suitable for forming quasars is not merely determined by gravitational parameters, as the hydro processes are also involved. Therefore, the identification of quasar-harboring halos should be given by both gravitational and hydro parameters. In other words, not all halos with certain $M$ and/or velocity dispersion $\\sigma_v$ harbor quasars, because certain hydro conditions must be satisfied. However, considering the hydro processes are local, it is reasonable to assume that the hydro conditions may not be modulated by the density inhomogeneities on scales much larger than the size of the halo $l$. In this case, the probability for a halo to have a quasar should be the same for all halos of the same kind, without depending on structures larger than $l$. Thus, the relative fractions of quasars with respect to the certain collapsed halos should be the same for all volumes larger than $l^3$. Consequently, when averaged on scale larger than $l$, the distribution of quasars $n_{qso}({\\bf r},z)$ at redshift $z$ should be proportional to that of the considered halos, $n_{halo}({\\bf r},z)$. Thus, all effects of the hydro processes can be absorbed into a ``normalization factor\" $A$, i.e. $n_{qso}({\\bf r},z)=An_{halo}({\\bf r},z)$, and $A$ is less dependent on $z$ than $n_{halo}({\\bf r},z)$. The bias of quasar distribution with respect to the mass distribution is then dominated by the bias of the selected halos with respect to the mass. Based on this analysis, quasar bias, at least on large scales, may also be only gravitational, depending on the gravitational parameters used for selecting the quasar-suitable halos. Accordingly, a possible model for quasar bias should at least satisfy the three conditions. 1. Gravitational environment given by the identified halos is consistent with the observed environment around quasars; 2. The abundance of quasars, $n_{qso}(z)$, at redshift $z$ is proportional to the number density of the identified halos, $n_{halo}(z)$ in a redshift-independent way, i.e. $n_{qso}(z)=An_{halo}$ where $A$ is a {\\it z-independent} constant, 3. The amplitude and $z$-evolution of the halo-halo correlation function are consistent with the observed correlation function of quasars. In this letter, we will show within the framework of the CDM cosmogonic theories that such a bias model can indeed be settled following the above-mentioned points. The details of the points 1, 2 and 3 will be discussed in the \\S 2, 3 and 4, respectively. ", "conclusions": "We showed that velocity-dispersion-selected halos are a possible mechanism for the bias of quasars. The majority of quasars at redshift $z\\sim 1 - 5$ formed in the environment of new born collapsed halos with 1-D velocity dispersion $\\sigma^{1d}_v \\sim 400 \\kms$. Both the harboring coefficient $f$ per halo and the lifetime of quasars are $z$-independent. With this bias model, the popular structure formation models, like SCDM and LCDM, can be fairly well reconciled with data of the abundance and correlations of quasars at $z \\geq 0.5$. It is interesting to point out that the velocity dispersion identified halos generally don't have the same mass. Eq.(2) shows that for a given $\\sigma^{1d}_v$, the redshifts the higher, the mass of the halos the smaller. This result has already been recognized in an earlier study, which shows that in order to fit with quasar abundance at high redshift, the mass of the halos has to be smaller than at the lower redshift (Bi \\& Fang 1997). With this model, one can predict that 1. The environment for quasars at redshifts from $z \\sim 1$ to 5 should be characterized by a velocity dispersion, $\\sigma^{1d}_v \\sim 400 \\kms$; 2. The amplitudes of quasar two-point correlation function at high redshifts don't significantly evolve with redshifts. In the paper, only the models of the SCDM and LCDM are considered. We can expected that with better data of quasars becoming available, the bias model of quasars will play more important role for discriminating among models of structure formations." }, "9804/astro-ph9804330_arXiv.txt": { "abstract": "s{We review present understanding of Galactic free--free emission and its possible importance to CMB fluctuation measurements. Current results, from both ``direct'' observations in the microwave band and from H$\\alpha$ studies, suggest that this foreground does not represent a serious obstacle to mapping the CMB; however, this is based on limited information and we emphasize the need for more exhaustive studies. We also present some preliminary results based on our recent H$% \\alpha$ observations near the South Pole CMB data sets. The fluctuation amplitude seen in H$\\alpha$ indicates that the detected CMB fluctuations are not significantly contaminated by free--free emission, at least if the diffuse gas is at a temperature of $T\\sim 10^4$ K.} ", "introduction": " ", "conclusions": "To summarize the current status of our understanding of Galactic free--free emission, we would say that although there is {\\em no indication} of fluctuations large enough to pose serious difficulties for CMB observations, the observational constraints remain weak. A critical interpretation of the results in the tables would be that the limit on large scale free--free fluctuations is the same order as the CMB amplitude on these same scales (at $~40$ GHz). On smaller scales, observations in $H\\alpha$ have not turned up any signs of large amplitude variations, but those based on high resolution spectrographs are few and cover only a small percentage of the sky. There does appear to be a dust/free--free correlation on all angular scales, but there is room, and perhaps tentative indications of, an equally important non--correlated component (question \\#1 posed in the introduction). And then there is the puzzeling result from Leitch et al. (1997), perhaps pointing to a hot phase of the ISM which could, due to lack of sensitivity, escape many of the present $H\\alpha$ limits (question \\#2 posed in the introduction). Obviously, CMB observations at higher frequencies, where much of the effort is now being concentrated, will suffer less from any possible free--free contamination, and the many of the next generation CMB experiments have a wide spectral coverage to aid the removal of foregrounds. Even given the above critical viewpoint, it would be a surprise to discover free--free emission presenting an important difficulty for all planned CMB experiments, at least in terms of measuring the variance, or power spectrum, from the early Universe. Foregrounds will, however, be much more important for higher order statistics looking for non--gaussian signatures. In such cases, the non--gaussian foregrounds will have to be removed to high precision. A sensitive, high spectral resolution H$\\alpha$ survey of the entire sky would be of great value in the context of the above considerations. We have also reported some results from our recent H$\\alpha$ observations taken with a Fabry--Perot system at La Silla. The data cover two bands along which the South Pole data sets show significant fluctuations in CMB brightness. The observed H$\\alpha$ fluctuations indicate that the CMB results in this region are not significantly contaminated by free--free emission (assuming a gas with $T_4=1$)." }, "9804/astro-ph9804218_arXiv.txt": { "abstract": "We present new spectroscopic and photometric time series observations of the $\\delta$ Scuti star FG~Vir. We detect the oscillations via changes in the equivalent widths of hydrogen and metal absorption lines. {}From the ratios between spectroscopic and photometric amplitudes, we assign $\\ell$ values to the eight strongest oscillation modes. In particular, we identify two radial modes ($\\ell =0$) and find that the main pulsation mode (147~$\\mu$Hz) has $\\ell =1$. One of the radial modes (at 140$\\mu$Hz) is the fundamental, implying that two modes with lower frequencies are {\\it g}-modes. For the radial modes, we compare frequencies with those calculated from a scaled $\\delta$~Scuti star model and derive a density $0.1645\\pm 0.0005\\,\\rho_{\\odot}$. We then obtain a distance of $84\\pm 3$\\,pc, in excellent agreement with the Hipparcos value. Finally, we suggest that a 3.5-day variability in all observables (equivalent widths and intensity) is caused by stellar rotation. ", "introduction": "\\label{intro} Oscillations in multi-periodic variables such as $\\delta$ Scuti, roAp and $\\beta$ Cephei stars have been observed extensively during the past 20 years. But even with high-quality data, it is still extremely difficult to identify which modes are being detected. Kjeldsen et al.\\ (\\cite{kbvf95}) used a new technique to detect solar-like oscillations in the bright G sub-giant $\\eta$ Boo through their effect on the equivalent widths of the Balmer lines. A subsequent discussion by Bedding et al.\\ (\\cite{bkrb96}) of the sensitivity of different observables to modes with different degree $\\ell$ suggested that one can determine the $\\ell$-value of a given mode by combining measurements of absorption-line equivalent widths with simultaneous photometric observations. To test this idea, we chose the bright and well-studied $\\delta$~Scuti star FG~Vir. FG Vir (HD 106384; $V=6.57$) is a multi-periodic $\\delta$~Scuti star. It has a main pulsation period close to 1.9 hours and shows a fairly complex oscillation spectrum. This star has been studied extensively during the last few years, resulting in the detection of at least 24 well-determined frequencies between 100 and 400~$\\mu$Hz, with amplitudes from 0.8 to 22 milli-magnitudes (mmag; Breger et al.\\ \\cite{bhn95}, \\cite{br98}). Because of its slow rotation (which reduces the complicating effects of rotational splitting) and the large number of detected frequencies (some of which are probably {\\it g} modes), FG Vir is one of the most promising candidates for performing asteroseismology on a $\\delta$ Scuti variable. Observations and models of this star have been presented by Dawson et al.\\ (\\cite{dbl95}), Breger et al.\\ (\\cite{bhn95}) and Guzik \\& Bradley (\\cite{gb95}). By choosing a star like FG Vir we have the advantage of knowing the frequencies in advance. We are therefore able to determine the oscillation amplitudes and phases with high precision. The aim of this paper is to identify the $\\ell$ values of the observed modes and to compare the oscillation frequencies with a pulsation model. A preliminary analysis of the observations presented in this paper was given by Viskum~et~al.~(\\cite{vdk97}), while results on radial velocity measurements were given by Viskum et al.\\ (\\cite{vb97}). ", "conclusions": "We have investigated a new technique to measure the oscillations in $\\delta$~Scuti stars via changes in the equivalent widths of absorption lines. An important advantage of this new technique is that only medium-dispersion spectra are needed, which makes the method suitable for small and medium-sized telescopes and for multi-site campaigns. Our main results are summarized below. \\begin{itemize} \\item Our detection of oscillations in FG~Vir from equivalent-width measurements of \\Ha, \\Hb\\ and Fe\\,{\\sc i} lines is an important confirmation of the method developed by Kjeldsen et al.\\ (\\cite{kbvf95}) to search for solar-like oscillations. \\item {}From the ratios between oscillation amplitudes measured in EW and the four Str\\\"omgren filters ({\\it uvby}), we have identified $\\ell$-values for eight modes in FG~Vir. \\item We suggest that the two lowest-frequency modes are {\\it g}-modes, while the strongest mode (147.2$\\mu$Hz) is a dipole mode. \\item {}By comparing the frequencies of radial modes with model calculations, we obtained a precise density and derived a distance that is in excellent agreement with the Hipparcos value. \\item We detected a long-period variation in the time series with a period of 3.5 days, which we propose is caused by rotation of the star. \\end{itemize}" }, "9804/astro-ph9804112_arXiv.txt": { "abstract": "We discuss a technique for mapping the synchrotron turnover frequency distribution using nearly simultaneous, multi--frequency VLBI observations. The limitations of the technique arising from limited spatial sampl\\-ing and frequency coverage are investigated. The errors caused by uneven spatial sampl\\-ing of typical multi--frequency VLBA datasets are estimated through numerical simulations, and are shown to be of the order of 10\\%, for pixels with the deconvolution ${\\rm SNR} \\sim 7$. The fitted spectral parameters are corrected for the errors due to limited frequency coverage of VLBI data. First results from mapping the turnover frequency distribution in \\object{3C\\,345} are presented. \\keywords {methods: data analysis -- methods: observational -- quasars: individual: \\object{3C\\,345}} ", "introduction": "} Information obtained with Very Long Baseline Interferometry (VLBI) about radio spectra of parsec--scale jets and their evolution can be crucial for distinguishing between various jet models. However, there are several aspects of VLBI which impede spectral studies of parsec--scale regions. The reliability of spectral information extracted from VLBI data depends on many factors including sampl\\-ing functions at different frequencies, alignment of the images, calibration and self--calibration errors, a narrow range of observing frequencies, and source variability. The influences of all these factors must be understood and, if possible, corrected for, in order to reconstruct the spectral properties of parsec--scale jets consistently. Radio emission from the parsec--scale jets is commonly described by the synchrotron radiation from a relativistic plasma (e.g. Pacholczyk 1970). The corresponding spectral shape, $S(\\nu) \\propto \\nu^{\\alpha}$, is characterized by the location of spectral maximum ($S_{\\rm m}$, $\\nu_{\\rm m}$) also called the turnover point, and by the two spectral indices, $\\alpha_{\\rm thick}$ (for frequencies $\\nu \\ll \\nu_{\\rm m}$) and $\\alpha_{\\rm thin}$ (for $\\nu \\gg \\nu_{\\rm m}$). In many kiloparsec--scale objects, spectral index distributions have been mapped, using observations made with scaled arrays. In such observations, the antenna configurations are selected at each frequency in a specific way such that the spatial sampl\\-ings of the resulting interferometric measurements are identical at all frequencies used for the observations. It is virtually impossible to use the scaled array technique for VLBI observations of parsec--scale jets made at different frequencies. The uneven spatial sampl\\-ings of VLBI data at different frequencies result in differences of the corresponding synthesized beams, and can ultimately lead to confusion and spurious features appearing in spectral index maps. In spectral index maps, the only available kind of information is the spectral slope between the two frequencies. While sufficient for many purposes, this information can be misleading in the situation when the frequency of thespectral maximum lies between the frequencies used for spectral index mapping. In the ranges of frequencies between 1.4 and 43\\,GHz, frequently used for VLBI observations, such a situation can be quite common. Using observations at three or more frequencies, it is possible to estimate the shape of the synchrotron spectrum, and derive the turnover frequency (frequency of spectral maximum). Information about the turnover frequency can help to avoid the confusion which is likely to occur in spectral index maps. The turnover frequency is sensitive to changes of physical conditions in the jet such as velocity, particle density, and magnetic field strength. This makes it an excellent tool for probing the physics of the jet in more detail than is allowed by analysis of the flux and spectral index properties of the jet. In this paper, we present a technique suitable for determining the turnover frequency distribution from multi--frequency VLBI data, and investigate its limitations and ranges of applicability. We discuss the advantages of using the Very Long Baseline Array\\footnote{The Very Long Baseline Array is operated by the National Radio Astronomy Observatory (NRAO)} (VLBA) for spectral imaging. A general approach to imaging of VLBA data from nearly simultaneous, snapshot--type observations at different frequencies is outlined in section~\\ref{sc:imaging}. The effects of limited sampl\\-ing and uneven {\\it uv}--coverages are discussed in section~\\ref{sc:spsens}. We provide analytical estimates of the sensitivity decrease, and use numerical simulations to evaluate the effect the uneven spatial sampl\\-ings have on the outcome of a comparison of VLBI images at different frequencies. Alignment of VLBI images is reviewed in section~\\ref{sc:imalign}. A method used for spectral fitting and determining the turnover frequency is described in section~\\ref{sc:fitting}. Spectral fitting in the case of limited frequency coverage is discussed in section~\\ref{sc:frcoverage}. The first results from the turnover frequency mapping are presented in section~\\ref{sc:algorythm}. ", "conclusions": "} In this paper, we have covered several methodological and scientific aspects of studying synchrotron spectrum of the parsec--scale regions in AGN. The main conclusions can be stated as follows: 1)~We have discussed a technique that can be used for mapping the turnover frequency distribution and obtaining spectral information from multi--frequency VLBA data. A feasibility study shows that multi--frequency VLBA observations can be used for spectral imaging and continuous spectral fitting. 2)~Multi--frequency VLBA observations made with up to 10 minute separations between the scans at each frequency can provide a satisfactory spatial sampl\\-ing and image sensitivity for sufficiently bright sources with intermediate ($\\sim 10$--15\\,mas) structures. The fractional errors from comparing the data at different frequencies should not exceed 10\\% for emission with SNR$\\ge 7$, in this case. 4)~A procedure for broadband synchrotron spectrum fitting has been introduced for mapping the distribution of spectral parameters of radio emission from parsec--scale jets. Corrections based on the local curvature of the fitted spectra are introduced, in order to compensate for the incomplete frequency coverage in cases where the true turnover frequency is outside of the range of observing frequencies. 5)~From a 4--frequency VLBA observation of \\object{3C\\,345}, the first map of the turnover frequency distribution are produced. The maps indicate possible locations of the relativistic channel and strong shock fronts inside the jet. The magnetic field distribution derived from the turnover frequency and flux distributions is consistent with the plane shocks existing in the immediate vicinity of the source core. The extended emission appears to have a very low turnover frequency for which the existing data do not warrant a good estimate, limiting the conclusions to deducing certain information from the gradients of the turnover frequency which are visible in the extended jet. The observed gradients are consistent with the patterns of velocity distribution and density gradients typical for Kelvin--Helmholtz instabilities propagating in a relativistic jet. A more detailed study, with observations made at lower frequencies, is required for making conclusive statements about the nature of the observed gradients of the turnover frequency." }, "9804/astro-ph9804262_arXiv.txt": { "abstract": " ", "introduction": "The EGRET experiment on board CGRO revealed the existence of a diffuse gamma-ray background (hereafter DGRB) at the level of $I_{DGRB} = 9.6 \\cdot 10^{-7} E_{GeV}^{-2.11 \\pm 0.05} ~ cm^{-2} s^{-1} sr^{-1} GeV^{-1}$ \\cite{owz94} in the energy range $0.03 \\div 10$ GeV. However, a recent reanalysis of the EGRET data \\cite{swz} found that the level of the DGRB is systematically lower by a factor $\\sim 20 \\%$ in the energy range $\\sim 0.1 \\div 4$ GeV. The DGRB is observed at high galactic latitudes $b > 10$ deg and such an evidence suggested an extragalactic origin for this diffuse background. Nonetheless, the specific origin of the DGRB is still under debate. In fact, the EGRET experiment \\cite{kan} has a poor angular resolution ($\\theta_{min} \\sim 1$ deg) so that it is hard to discriminate among different origins of this extragalactic background. Specifically, it is still difficult to discriminate between a purely diffuse nature of the DGRB (see \\eg \\cite{ca}) and the option of a DGRB made by a superposition of unresolved, discrete sources. The large number of identified AGNs and flat spectrum radio quasars (hereafter FSRQ) in the EGRET sky (\\cite{fi96} \\cite{has} \\cite{mat}) suggested that most of the DGRB can be produced by a non-resolved population of AGNs, the actual fraction of the DGRB produced by FSRQ and BLLacs being in the range $\\sim 40 \\div 95 \\%$ (see \\eg \\cite{com}). Separately, it has been evaluated that a fraction $\\sim 42 \\div 97 \\%$ of the DGRB could be ascribed to blazars \\cite{pad}. However, the flatness of the spectrum of the DGRB seems to favour the possibility that BLLacs could be the major contributors to the DGRB of extragalactic origin \\cite{pohl}. Erlykin \\ea \\cite{erl96} reviewed the various AGN contributions and quoted that the fraction of the DGRB produced by the observed AGNs is $\\sim 65 \\%$. The DGRB fractions previously reported may be subject to a revision ($\\sim 25 \\%$ increase) if the recent reanalysis \\cite{swz} of the EGRET data is adopted. On account of the large theoretical uncertainties and of the present observational precision of the EGRET detectors, it is still hard to discriminate among the different proposed possibilities, even though a fluctuation analysis of the EGRET data should give more precise indications on the nature and origin of the DGRB. Beside the discrete, unresolved source case pictured for the origin of the DGRB, there have been some pioneering works \\cite{hw} \\cite{ds} \\cite{bbp} \\cite{volk} suggesting that a relevant fraction of the DGRB could be produced by {\\it extended} sources through hadronic collisions of cosmic ray (hereafter CR) protons interacting with the protons of the Inter Galactic Medium (hereafter IGM) which is abundantly present within galaxy clusters (see \\cite{sa88} for a review). In this alternative picture, the CR's are assumed to be produced within clusters (we will discuss in Sect.4 some of the possible sources) where also a population of protons and electrons is residing in the form of a hot (with temperatures $T \\sim 10^7 \\div 10^8$ K), tenuous (with electron number densities $n_e \\sim 10^{-3}~cm^{-3}$), chemically enriched and massive (with mass fractions $M_{IGM}/M \\sim 0.05 \\div 0.3$) plasma: the IGM. The proposed mechanism has an essential ingredient in the confinement of the CR's within clusters where they are produced; this point, already realized by some authors \\cite{bbp} \\cite{volk}, is responsible for the net increase in the probability of interaction per proton with respect to the case of a straight line propagation. The increase factor can be estimated to be $\\sim c t_{cl} /R_{cl}\\simgt 600$, where $t_{cl} \\simlt H_0^{-1}$ is the age of the cluster and $R_{cl}$ is its size. Cosmic rays produced within a cluster during all its lifetime can thus produce gamma rays through the production and the subsequent decay of neutral pions: \\begin{equation} p+p\\to \\pi^0+X~,~~~~~~~~~~~~~~\\pi^0\\to \\gamma+\\gamma. \\end{equation} Note that in the same interactions, charged pions are also produced, which determine a neutrino emission through the following channels: \\begin{equation} p+p\\to\\pi^{\\pm}+X, ~~~~~\\pi^{\\pm}\\to \\mu^{\\pm} \\nu_{\\mu}(\\bar{\\nu}_{\\mu}), ~~~~~\\mu^{\\pm}\\to e^{\\pm} + \\bar{\\nu}_{\\mu}(\\nu_{\\mu}) + \\nu_e (\\bar{\\nu}_e) ~. \\end{equation} We will also discuss the relevance of these last processes in Section 7 below. Using the gamma ray production from clusters of galaxies according to eq. (1), Houston \\ea \\cite{hw} suggested that the total extragalactic gamma ray intensity detected above $35$ MeV \\cite{ft82}, $I_{\\gamma} \\approx 5.5 \\cdot 10^{-5}~ cm^{-2} s^{-1} sr^{-1}$, could be ascribed, for a large fraction, to galaxy clusters. They predicted a level $I_{\\gamma} \\approx 5 \\cdot 10^{-5} cm^{-2} s^{-1} sr^{-1}$ at energies above $35$ MeV, assuming an observed local cluster space density, $n_{cl} \\approx 7.3 \\cdot 10^{-5} Mpc^{-3}$, integrated out to the Hubble radius, $R_H=6 \\cdot 10^3$ Mpc, and neglecting any cosmological effect. More recently, Dar \\& Shaviv (hereafter DS \\cite{ds}) reanalyzed the problem in the light of the EGRET data \\cite{owz94} and calculated the contribution to the DGRB from CR interactions in the intracluster gas, under the assumption that the energy density of CR's in clusters is the same as in our own galaxy (universality). With this assumption, Dar \\& Shaviv \\cite{ds} predicted a level $I_{\\gamma}(> 100~ MeV) \\approx 1.2 \\cdot 10^{-5}$ photons cm$^{-2}$ s$^{-1}$ sr$^{-1}$, which could explain the whole amount of the DGRB of extragalactic origin. In a following paper, Berezinsky, Blasi \\& Ptuskin (hereafter BBP \\cite{bbp}) relaxed the {\\it ad hoc} assumption of universality, and estimated the CR energy density in clusters due to various possible sources of CR, using the condition of diffusive confinement of CRs. In their approach, BBP \\cite{bbp} showed that it is impossible to fulfill the universality condition with the usual CR sources in clusters, emphasizing that the DGRB due to the CR interactions in clusters should be a small fraction of the total diffuse flux observed by EGRET. This conclusion was reached by the previous authors under the hypothesis that a large fraction of the baryons in the universe is contained inside clusters of galaxies (BBP considered that clusters are a fair sample of the baryons in the universe \\cite{wf91}, \\cite{wetal93}, \\cite{wf95}) assumed to have a homogeneous inner distribution of gas, $n_e =const$ (here $n_e$ is the IGM electron number density). Because of these assumptions, their results depend only on overall cosmological parameters like the baryon fraction in the universe $\\Omega_b$, and on the cluster size. Dar \\& Shaviv \\cite{ds} also predicted the gamma ray fluxes from a few nearby clusters (Coma, Perseus and Virgo): for these three clusters they found $\\gamma$-ray fluxes in the range $F_{\\gamma}(>100 ~MeV) \\approx 5 \\div 20 \\cdot 10^{-8} cm^{-2} s^{-1}$. In particular, the value which they predicted for A1656 (Coma), $F_{\\gamma}(>100 ~MeV) \\approx 5 \\cdot 10^{-8} cm^{-2} s^{-1}$, is close to - or slightly higher than - the upper limits given by EGRET for this source. Similar results were obtained for these clusters by Ensslin \\ea \\cite{ensslin96} assuming a population of CRs from radio sources located within galaxy clusters in almost equipartition with the IGM thermal energy. We stress here that in all the previous works a uniform IGM density profile was assumed. Moreover, the cluster population was not assumed to evolve with cosmic time, and the same working hypothesis of no-evolution was assumed for the IGM content of each cluster. However, X-ray studies of galaxy clusters, have shown that these cosmic structures are indeed well structured, having a gas density profile $n(r) \\propto [1+(r/r_c)^2]^{- 3 \\beta /2}$, with core radii $r_c \\approx 0.1 \\div 0.3 \\hmpc$ and $\\beta \\approx 0.6 \\div 0.8$ (see \\eg \\cite{jf92}; see also \\cite{sa88} and references therein). Beside this, the IGM is indeed evolving as indicated by its sensitive metal enrichment, $Fe/H \\sim 0.2 \\div 0.5$ (in solar units, see \\eg \\cite{e90}, \\cite{ar}), shown even for the brighter clusters observable at redshifts $z \\sim 0.5$ \\cite{lm97}. Nonetheless, there is also an increasing debate on the possible evolution of the X-ray luminosity function observed out to $z \\simlt 0.5$ with the EINSTEIN \\cite{gio90} \\cite{h92} and ROSAT satellites \\cite{ebe97} \\cite{nich} and on the possible evolution of the cluster temperature function \\cite{cmv} \\cite{eke98} \\cite {vl98}. If an evolution is present in the cluster population this can be, in fact, understood as a result of two competing effects: \\newline {\\it i)} a luminosity evolution, where the cluster X-ray luminosity, $L \\propto n^2 T^{1/2} R^3$ (mainly due to thermal bremsstrahlung), changes with redshift due to variations in the gas mass density, $n \\propto f_g \\rho_{cl}$ (where the cluster gas mass is taken to be a fraction $f_g \\equiv M_{gas}/M$ of the total cluster mass), and/or changes in the IGM temperature $T$ at fixed mass, $M \\propto \\rho_{cl} R^3$ (here $\\rho_{cl}$ is the cluster total mass density); \\newline {\\it ii)} a change in redshift of the number density, $N(M,z)$, (usually referred to as mass function, hereafter MF) of clusters that are found to be collapsed (or virialized) in the mass range $M, M+dM$ at redshift $z$. Detailed studies of cluster evolution in X-rays (see \\eg \\cite{cv}, \\cite{ob}, \\cite{cmv}) considered in fact that a combination of the previous mechanisms is responsible for the actual cluster evolution when they fit the available data (see \\cite{cv} for a detailed discussion). In this paper we predict the amount of high energy, non-thermal, gamma-ray emission from galaxy clusters using detailed modelling of the realistic cluster structure, as well as viable modelling for the evolution of the IGM and of the cluster MF. Based on these phenomenological cluster models, we predict the amount of DGRB that can be produced in the viable cosmological models: here we consider flat and low-density (open or vacuum-dominated) CDM models as well as mixed Dark Matter models with a fraction $\\Omega_{\\nu} \\approx 0.3$ of the total density of the universe in form of massive neutrinos. We use $h=H_0/100$ km s$^{-1}$ Mpc$^{-1}$ throughout the paper unless otherwise specified. The plan of the paper is the following. In Sect.2 we briefly summarize the cluster formation hystory in hierarchical scenarios for structure formation. In Sect.3 we describe a model for the production of diffuse gamma-ray emission due to the interaction of CR's with the target protons present in the extended, diffuse IGM. We consider in Sect. 4 different CR sources that can be found in connection with galaxy clusters. We discuss in Sect.5 the correlation between the extended gamma-ray emission and the much better known thermal X-ray emission coming from the IGM. Based on these properties, we construct a list of predicted $\\gamma$-ray fluxes for a compilation of X-ray clusters with detailed informations on their IGM structure, IGM temperatures and X-ray fluxes. In Sect. 6 we present predictions for the amount of DGRB produced by galaxy clusters in different cosmological scenarios. We briefly discuss in Sect.7 the extended neutrino fluxes emerging from these objects and their contribution to a possibly detectable diffuse neutrino background (hereafter DNB). Finally, in Sect.8 we discuss our results in the light of the current limits obtained from EGRET and in the light of the future experiments for gamma-ray and neutrino astronomy. ", "conclusions": "In this paper we presented a detailed study of the diffuse emission of $\\gamma$-rays and neutrinos from clusters of galaxies. Using realistic modelling of the cluster structure, of their formation history and of their evolution with cosmic time, we found that galaxy clusters can provide $\\simlt 1\\%$ of the DGRB measured by EGRET (in the first release by OWZ \\cite{owz94}). Our estimate of $I_{\\gamma}$ is quite independent on the geometry of the universe, on the assumed cosmological model and on the amount of IGM evolution, because most of their contribution to the DGRB comes from nearby, $z \\simlt 0.2$, clusters. In fact, at these redshifts the effects of curvature do not take place strongly in changing the perturbation growth factor, ${\\cal D}(z,\\Omega_0)$, (normalized at the present epoch), the difference in cluster evolution are small when the different models are normalized to the local abundance of clusters observed in X-rays and the available amount of IGM evolution - even if considered to be quite strong, $f_g \\propto (1+z)^{-1 \\div -2}$ - can provide only small variations to the cluster $\\gamma$-ray luminosities, as $L_{\\gamma} \\propto f_g$ (see eq.18). On account of all these aspects, we consider that our results for the contribution of galaxy clusters to the DGRB are quite robust. Our approach differs substantially from the previous ones in several (among others) aspects: \\par\\noindent {\\it i)} we considered - differently from all the previous approaches - a self-consistent approach to the formation of clusters following the spherical collapse model \\cite{peeb80} complemented with a realistic IGM density profile, consistent with the most recent determinations from X-ray observations. This fact has important effects on the CR confinement within cluster cores and hence on the relative $\\gamma$-ray and neutrino emission rates; \\par\\noindent {\\it ii)} we considered (as BBP did) here an energy dependent diffusion coefficient which results in a very general picture of the CR confinement within cluster cores; \\par\\noindent {\\it iii)} we also considered here - at variance with the previous approaches - the effects of a possible evolution in the cluster IGM content. This is consistent with the present indications of a variation in the IGM content from groups to rich clusters in the local frame and with the X-ray, shock (or entropy) induced, luminosity evolution observed from numerical simulations \\cite{tm97} and predicted in analytical models (both shock and entropy models) for the evolution of X-ray clusters \\cite{cola97} \\cite{b97}; \\par\\noindent {\\it iv)} we use the PS cluster MF that was found to be consistent with N-body simulations over a large dynamical range and up to $z \\simgt 2 $ \\cite{eke96}. We normalized it to the local abundance of clusters detected in X-rays. In the previous approaches average values for the overall cluster abundance, $n_{cl} \\approx 4\\div 7 \\cdot 10^{-5}~ Mpc^{-3}$, were used \\cite{hw} \\cite{bbp} without considering the effect of an evolving cluster mass function; \\par\\noindent {\\it v)} using a self-consistent modelling of the IGM we found an analytical correlation between $\\gamma$-ray and $X$-ray emission for clusters. The predicted ratio $F_{\\gamma} / F_X \\propto f_g^{-1} r_c^{-1} T^{-1/2}$ [see eq.(32)] provides a behaviour of the $F_{\\gamma}-F_X$ relation different from that obtained by Ensslin \\ea \\cite{ensslin96}, $F_{\\gamma} / F_X \\propto T^{1/2}$, because we did not assume any (partial) equipartition between IGM thermal energy and relativistic jet particles. Our correlation results only from the basic electromagnetic and hadronic emission mechanisms in which the IGM protons are the targets for both the X-ray thermal bremsstrahlung emission and for the $p p$ collisions responsible for $\\gamma$-rays. \\par\\noindent {\\it vi)} using such a correlation we derived a sample of nearby clusters with predicted $\\gamma$-ray fluxes observable with the next generation $\\gamma$-ray telescopes. Incidentally, we found a $\\gamma$-ray flux for Coma, $F_{\\gamma}(>100 MeV)\\approx 8.5 \\cdot 10^{-9}~ photons$ $s^{-1} ~cm^{-2}$ which is consistent with the EGRET upper limits for this cluster (previous specific predictions \\cite{ds} \\cite{ensslin96} seem to exceed the EGRET upper limit). Our numerical results for $I_{\\gamma,\\nu}$ are in reasonable agreement with those obtained by BBP, even though based on a quite different description of the cluster structure and evolution. This agreement is due to the fact that BBP considered a constant comoving density of clusters, $n_{cl} \\sim 5 \\cdot 10^{-5}$ Mpc$^{-3}$, assumed to be a fair sample of the baryons in the universe, and containing a fraction $\\Omega_b\\approx 0.5 \\Omega_{BBN}$ (where $\\Omega_{BBN}$ is the value of the baryon density predicted by Big Bang Nucleosynthesis). Under these assumptions, BBP obtained a value for $I_{\\gamma}$ higher by a factor $\\sim 3\\div 4$ with respect to our result, based on values $f_g \\sim 0.1$ (see Sect.2). Our refined calculations show why their assumption of considering $n_{cl} \\sim const$ was reasonable: the clusters effectively contributing to the DGRB are located at $z \\simlt 0.2$ (see Fig.5), where the effects of evolution do not have room to take definitely place (see Fig.7). Because of the inherent uncertainties in the predictions of quantities whose calculation involve to set the values of parameters which are not known precisely, we also estimated the range spanned by $I_{\\gamma, \\nu}$ for the combination of parameters allowed by the present observational ranges. In fact, the description of the cluster structure and evolution that we used in our analytic approach consider only ensemble averaged quantities. But we observe a whole distribution of the real cluster properties with respect to the average cluster moulding. Some amount of variance is needed to be considered in cluster modelling in order to ensure the predictive power of the viable models for structure formation. To explore the role of the uncertainties in the relevant quantities we considered several sources of uncertainties (see Sect.7). From an inspection of Figs. 9 and 10 we note that the effects of the possible theoretical uncertainties in the description of the cluster and IGM evolution could change the predicted contributions for the DGRB and for the DNB by a factor $\\simlt 3$, setting the maximal level of $I_{\\gamma}$ to a few $\\%$ of the EGRET value. The DGRB seems to be mostly produced by AGNs (FSRQ and/or BLLacs) and/or blazars (we take here an estimate of $\\sim 60 \\div 65 \\%$ \\cite{erl96} of the EGRET diffuse flux \\cite{owz94}). Diffuse $\\gamma$-ray emission could be observed also from normal galaxies yielding a contribution $\\sim 5 \\%$ \\cite{erl96}. When added to the $\\sim 10 \\div 15 \\%$ of the DGRB contributed by their high-$E$ photons interacting with other existing backgrounds (\\eg IR, CMB \\cite{erl96}) one gets only $\\sim 15 \\div 20 \\%$ of the DGRB left for truly diffuse or extended sources. Of this amount, a fraction of the diffuse $\\gamma$-ray flux $\\sim 3 \\div 5 \\%$ is predicted \\cite{ww} to originate from decaying topological defects (see \\cite{sigl}) and interactions of UHE particles with the CMB. Note, however, that the amount and the spectral distribution of this possible diffuse background depend sensitively on the amplitude of the primordial magnetic field on scales larger than supercluster sizes. So, according to the previous estimates, the presence of all these sources of diffuse $\\gamma$-ray emission (even though partially model dependent) determines an upper limit to the contribution of extended extragalactic sources to the DGRB, that is $\\sim 10 \\div 22 \\%$ of the OWZ EGRET level \\cite{owz94}. This sets rather weak constraints on the level of CR production in clusters and hence on the presence and activity of AGNs in clusters or on the formation and efficiency of accretion shocks around clusters. However, if we consider the revised level of the DGRB as derived by SWZ \\cite{swz}, then the previous upper limit reduces to $\\simlt 2 \\%$. Our predictions of the DGRB contributed by galaxy clusters $I_{\\gamma,cl} / I_{EGRET} \\sim 0.005 \\div 0.02$ is perfectly compatible with the presence of both a population of evolving FSRQ and AGNs dominating the $\\gamma$-ray sky and with the presence of truly diffuse backgrounds like those previously discussed. Note, however, that the major source of uncertainty in the level of the extragalactic DGRB comes from the contribution of the AGNs. A fluctuation analysis of the EGRET data is needed to have more definite indications on the level of the DGRB contributed by discrete sources. If, on the other hand, CR acceleration will be found to be relevant in clusters (yielding $L_{CR}$ substantially larger than $ 10^{44}$ erg/s), then the predicted level of DGRB produced by galaxy clusters can set interesting limits to the space density and evolution of $\\gamma$-ray AGNs. The sensitivities and angular resolutions achievable by the next generation gamma-ray (INTEGRAL, GLAST, AMS) and neutrino (see \\cite{halz} for a review) detectors will be able to shed a new light on the high energy phenomena occurring in large scale structures. \\vskip 1.5truecm \\newline {\\bf Acknowledgements} We thank the Referee for useful comments and suggestions which improved substantially the presentation of the paper. We also aknowledge interesting and stimulating discussions with V.S. Berezinsky during a recent visit of S.C. at the LNGS. S.C. aknowledges also interesting discussions with G. Kanbach, A. Dar and F. Halzen, among others, at the 1997 Moriond Meeting {\\it High Energy Phenomena in the Universe}. The research of P.B. is funded by a INFN PostDoctoral Fellowship at the University of Chicago. \\newpage" }, "9804/astro-ph9804054_arXiv.txt": { "abstract": "We present the results of an {\\asca} observation of the LINER NGC~4579. A point-like X-ray source is detected at the nucleus with a 2--10~keV luminosity of $1.5\\times10^{41}$~{\\eps}assuming a distance of 16.8~Mpc. The X-ray spectrum is represented by a combination of a power-law with a photon index of $\\sim$1.7 and soft thermal component with $kT\\sim$0.9~keV. An iron K emission line is detected at $6.73\\pm0.13$~keV (rest frame) with an equivalent width of $490^{+180}_{-190}$~eV and is statistically significant at more than 99.9\\% confidence. The line center energy is consistent with Helium-like iron and is significantly higher than 6.4~keV which is expected from fluorescence by \"cold\" (or a lower ionization state of) iron. The iron line profile shows no significant red tail in contrast to Seyfert 1 galaxies although the statistics are limited. The line center energy, equivalent width, and profile are consistent with an origin in an ionized accretion disk. However the large mass accretion rate necessary to ionize the accretion disk is not consistent with the observed luminosity and normal accretion models. ", "introduction": "Recent optical spectroscopic surveys have shown that there are many active galactic nuclei (AGNs) in nearby galaxies and about 40\\% of bright galaxies are classified as Seyfert galaxies or LINERs (Low Ionization Nuclear Emission-line Regions; \\cite{hec80}) (Ho, Filippenko, \\& Sargent 1997a). The luminosity of these objects are rather low compared to previously known AGN with a median value of the H$\\alpha$ luminosity being only $2\\times10^{39}$ {\\eps} in the sample of Ho et al (1997a). Such objects (low luminosity AGNs; hereafter LLAGNs) are important for investigating the physics of AGN under an extreme condition, i.e. very low luminosity. X-ray observations probe the innermost regions of AGNs and specifically the iron K line provides information on the ionization state, density, and motion of matter very close to the central energy source. {\\asca} observations of Seyfert 1 galaxies revealed that as a class these objects have a broad iron K line with a profile skewed to lower energies, thought to be caused by the reprocessing of the continuum by a relativistic accretion disk (e.g., \\cite{tan95}, Nandra et al. 1997a). The center energy of the iron line from Seyfert 1 galaxies is consistent with 6.4~keV, which is expected from fluorescence by neutral or lower ionization states of ($<$\\ion{Fe}{16}) iron in a disk with an inclination of $<$ 30 degrees. In the Seyfert 1.9 galaxy IRAS ~18325--5926, a higher peak energy of iron emission is seen, which is compatible with a highly-inclined disk ($i=40-50^{\\circ}$) origin (\\cite{iwa96}). Highly-ionized iron emission lines are detected from several radio-quiet quasars e.g. E1821+643 (\\cite{kii91}, \\cite{yam97}) and PG~1116+215 (\\cite{nan96}). Nandra et al. (1997b) studied the luminosity dependence of the iron line profile in a large sample of AGN and found that the center energy increases and the red-tail becomes weaker with increasing luminosity. They attributed such behavior to an increasing ionization of the accretion disk with increasing luminosity. Thus X-ray measurements of iron emission lines are powerful diagnostic tools of matter in the vicinity of the nucleus. There are only a few observations of iron emission lines from low luminosity AGNs ( {\\LX} (2--10~keV) $\\sim 10^{40}-10^{41}$~{\\eps}). M81 (NGC~3031) with an X-ray luminosity of {\\LX} (2--10~keV) $\\sim 2\\times10^{40}$~{\\eps} shows a broad iron line centered at $\\sim6.7$~keV with an equivalent width of $\\sim 200$~eV. This line center energy is significantly higher than Seyfert 1 galaxies and similar to luminous quasars. An iron line at 6.4~keV with an equivalent width of $\\sim300$~eV is detected from the low luminosity Seyfert 1 galaxy NGC~5033 ({\\LX} (2--10~keV) = $2\\times10^{41}$~{\\eps}, \\cite{te98b}) but only an upper limit on the equivalent width of {\\simlt} 300~eV is obtained for NGC~1097 ({\\LX} (2--10~keV) = $1\\times10^{41}$~{\\eps}, Iyomoto et al. 1996). Although strong iron emission lines are also detected from M51 (= NGC~5194, Terashima et al. 1998a), NGC~1365 and NGC~1386 (Iyomoto et al. 1997), the iron lines in these objects are interpreted as being caused by reprocessed emission from an obscuring tori and/or extended ionized scatterer outside of our line of sight, i.e., these nuclei are heavily obscured. Thus, at present, the number of LLAGNs with small intrinsic absorption from which iron lines are detected is rather limited. NGC~4579 (M58) is a Sab galaxy in the Virgo cluster of galaxies and classified as a LINER or Seyfert 1.9 galaxy based on the optical emission lines (\\cite{ho97a}, \\cite{kee83}, \\cite{sta82}) and the broad H$\\alpha$ component, detected with a FWHM $\\sim 2300$~km s$^{-1}$ (\\cite{ho97b}). There exists a flat-spectrum radio core (\\cite{hum87}). An {\\Einstein} HRI observation showed the presence of an unresolved X-ray nucleus and the X-ray flux was measured to be {\\FX} = $7.9\\times10^{-12}$~{\\eps} cm$^{-1}$ in the 0.2--4.0~keV band with the {\\Einstein} IPC (\\cite{fab92}, Halpern \\& Steiner 1983) which corresponds to the X-ray luminosity of $2.7\\times10^{41}$~{\\eps} (we assume a distance of 16.8~Mpc (\\cite{tul88}) throughout this paper). These facts indicate the presence of a LLAGN in this galaxy. A recent ultraviolet imaging observation by {\\it Hubble Space Telescope} ({\\HST}) Faint Object Camera (FOC) detected a point source at the nucleus (\\cite{mao95}). Its UV spectra were taken by {\\HST} Faint Object Spectrograph (FOS) and a featureless UV continuum is detected as well as various emission lines. Comparison of the FOC and FOS data also indicate a factor of 3.3 decrease of UV flux in 19 months. The narrow UV emission lines are incompatible with shock excitation model and a photoionization model is preferred (Barth et al. 1996). Several broad UV emission lines are also detected. These UV results provide further support for the presence of a LLAGN in NGC~4579. On the other hand, \\cite{mao98} estimated the ionizing photon number by extrapolating the UV luminosity at 1300 A towards higher energies and argued that the observed UV continuum is not sufficient to explain the H$\\alpha$ luminosity. They also suggest that emission from AGNs is most prominent at energies higher than the UV. Measurements of an X-ray flux and continuum slope provide information on the ionization source in this LINER. In this paper we report the detection of an Iron K emission line centered at 6.7~keV and discuss X-ray properties of the LLAGN in NGC~4579 and origin of the iron emission line. ", "conclusions": "\\subsection{X-ray emission from a low luminosity AGN} We obtained X-ray images and spectra in the 0.5--10~keV band and a point-like X-ray source with a photon index of $\\Gamma = 1.72\\pm0.05$ is detected. An iron line is also detected at 6.7~keV. In the soft energy band, a broad line like feature identified with iron-L line complex indicates the presence of thin-thermal plasmas of temperature $kT\\sim 0.9$~keV. The X-ray luminosity ($1.5\\times 10^{41}~${\\eps} in 2--10~keV ) is 1--3 orders of magnitude smaller than typical Seyfert galaxies and falls in the classes of LINERs and \"low luminosity\" Seyfert galaxies (Serlemitsos et al. 1996, Iyomoto et al. 1996, Ishisaki et al. 1996, Terashima et al. 1998b). In normal spiral galaxies, the X-ray emission is dominated by discrete sources, specifically low mass X-ray binaries (LMXBs) (e.g. \\cite{fa89}, Makishima et al. 1989). The X-ray luminosity from LMXBs are roughly proportional to B-band luminosity {\\LB} and their X-ray spectra can be approximated by a thermal bremsstrahlung of a temperature of several keV. The {\\asca} X-ray spectrum of NGC~4579 is also fitted by $kT\\sim8$~keV thermal plasma model. However the strong iron line at 6.7~keV is not compatible with the X-ray spectra of LMXBs, since the equivalent width of iron emission lines from LMXBs are small (several tens of eV, \\cite{hir87}). Additionally, the {\\LX}/{\\LB} value $1.3\\times10^{-3}$ is more than an order of magnitude higher than normal spiral galaxies (e.g. \\LX /\\LB =$3.5\\times10^{-5}$ for M31; \\cite{mak89}). Additionally an upper limit on the size of an archival {\\rosat} PSPC image is 14\" (Gaussian $\\sigma$), which corresponds to 1.1 kpc at 16.8 Mpc. This upper limit is significantly lower than the size of the galaxy. Therefore we conclude that contribution from LMXBs to X-ray emission of NGC~4579 is negligible. Hot plasmas with temperatures on the order of $\\sim10$ keV are present in the Galactic center region and their X-ray spectra show prominent, ionized iron K emission (e.g. Koyama et al. 1996). The X-ray spectral shape of NGC 4579 in the hard X-ray band is similar to such hot gas. However, the X-ray luminosity of NGC 4579 is three orders of magnitude higher than the Galactic ridge emission ({\\LX}$\\sim2\\times10^{38}$ {\\eps}; Kaneda et al. 1997, Yamasaki 1996, Warwick et al. 1985). Starburst galaxies also show a hard spectral component with a temperature of $\\sim$10 keV and their X-ray luminosities are around $10^{40}$ {\\eps} (e.g. $3.4\\times10^{40}$ \\eps in 2--10 keV for M82; Ptak et al. 1997). However the starburst activity in NGC 4579 is weaker than M82, since the far-infrared luminosity of NGC 4579 is about an order of magnitude lower than that of M82. Furthermore, starburst galaxies show weak or no iron emission contrary to NGC 4579. Therefore hot plasma is unlikely as the origin of the hard component and iron emission line in NGC 4579 and we conclude that the AGN emission dominates the {\\it ASCA} spectra and that other components such as a hot gas contribution is small, if any. We note that errors in background subtraction of the Virgo cluster hot gas do not affect the detection of the iron emission line at 6.7 keV, since the cluster gas is very dim in this region and temperature is low ($kT\\sim2$ keV; Matsumoto 1998, \\cite{boh94}). Actually no significant iron emission is detected from the GIS field around NGC 4579. If the primary ionizing mechanism of LINER optical emission lines in this galaxy is photoionization by a LLAGN, {\\LX}/{\\LHa} might be expected to be similar to Seyfert 1 galaxies, for which there is a good positive correlation between {\\LX} and {\\LHa} (e.g. Ward et al. 1988, \\cite{kor95}, Serlemitsos et al. 1996). Using the H$\\alpha$ luminosity of broad plus narrow component {\\LHa} = $5.9\\times10^{39}$~{\\eps} (\\cite{ho97b}) and the observed X-ray luminosity in the 2--10~keV band, we obtain {\\LX}/{\\LHa} $\\approx$ 26 for NGC~4579. This value is in excellent agreement with those of Seyfert 1 galaxies (\\cite{war88}) and strongly supports a low luminosity AGN as the ionizing source of the LINER in NGC~4579. Less luminous Seyfert 1 galaxies tend to show rapid and large amplitude variability (Mushotzky, Done, \\& Pounds 1993 and references therein). However NGC 4579 shows no significant short term variability. Lack of variability on short time scales seems to be a common property of LLAGNs (\\cite{mus92}, \\cite{pet93}), for example the LLAGN in NGC 1097 (Iyomoto et al. 1996) and NGC 3998 (\\cite{awa92}) also show no significant variability on timescales less than a day. Direct comparison of {\\rosat} PSPC and {\\asca} flux in the 0.5--2 keV band show a factor of two increase in $\\sim$3.5 years. The X-ray spectral slope $\\Gamma = 1.72\\pm0.05$ is identical to the average value found for hard X-ray selected Seyfert 1 galaxies (Mushotzky et al. 1993) but the luminosity is lower than that of any Seyfert 1 galaxy but NGC 4051. Based on the FW0I (full width at 0 intensity) of a broad emission line and an estimate of the size of the broad line region, mass of the central black hole is roughly estimated to be $M_{\\bullet} \\sim 4\\times10^6M_{\\odot}$. Then the Eddington ratio $L/L_{\\rm Edd}$ is $\\sim 10^{-3}$ for the observed luminosity of $\\sim5\\times10^{41}$~{\\eps} (Barth et al. 1996), although their blackhole mass estimation is crude. Therefore the X-ray spectral slope does not seem to be drastically changed even at a very low Eddington ratio. This is also true for M81, for which $L/L_{\\rm Edd}$ is estimated to be $\\sim (2-10)\\times 10^{-4}$ (\\cite{ho96}) and the photon index is $1.85\\pm0.04$ (\\cite{ish96}). Soft thermal emission of $kT\\sim0.5-1$~keV is often observed from low luminosity AGNs (Terashima 1997, \\cite{pta97}, Serlemitsos et al. 1996). In some cases, such emission is associated with starburst activity (e.g. \\cite{iyo96}, \\cite{te98a}). Since the far-infrared luminosity of NGC~4579 is $1.5\\times10^{43}$~{\\eps} some star formation activity may be present which may explain the thermal emission. The soft thermal X-ray to far infrared luminosity ratio {\\LX}/{\\LFIR} = $6\\times10^{-4} - 1.1\\times10^{-3}$ is consistent with starburst galaxies (e.g. \\cite{dav92}) within the scatter. \\subsection{Iron-K line} A marginally broad ($\\sigma \\approx 0.17$~keV) iron emission line is clearly detected at $6.73^{+0.13}_{-0.12}$~keV and the equivalent width is $490^{+180}_{-190}$~eV for the broad Gaussian model fit. The line center energy is significantly higher than 6.4~keV, which is typically observed from Seyfert 1 galaxies, and consistent with He-like iron. A similar broad iron line centered at $\\sim 6.7$~keV is detected from the low luminosity Seyfert galaxy M81 (Ishisaki et al. 1996, Serlemitsos et al. 1996). The line can also be represented by line blending of neutral, He-like, and H-like iron and dominated by He-like iron. The disk-line profile (Fabian et al. 1989) is probably inconsistent with the data for 6.4~keV or 6.7~keV intrinsic line energy because of following reasons. The $\\chi^2$ value is worse than a single broad Gaussian fit and systematic residuals remain in the disk-line fit, since a significant red tail is not clearly seen in the data. Furthermore, the disk-line model provides the very large equivalent width $\\sim900$~eV, which is about 4 times larger than the results of the disk-line fit to Seyfert 1 galaxies ($<$EW$>=(230\\pm60)$ ~eV, \\cite{na97a}). Therefore our data prefer a symmetric Gaussian-shape profile with intrinsic line center energy of 6.7~keV (He-like) rather than 6.4~keV ($<$\\ion{Fe}{16}). Thus the ionization state of the iron line emitter may be different from that of higher luminosity Seyfert 1 galaxies in at least some LLAGNs (NGC 4579 and M81) Strong ionized iron emission lines are observed in heavily obscured Seyfert 2 galaxies (NGC 1068, \\cite{uen94}, \\cite{iwa97}; NGC 1365, \\cite{iyo97}; see also \\cite{tu97a}, \\cite{tu97b}). In these objects continuum emission from the nucleus is completely blocked and only scattered radiation is observed. Ionized iron lines are interpreted as originating from a photoionized scattering medium. If the continuum of NGC~4579 is scattered radiation, then the observed X-ray luminosity is only a fraction of its intrinsic luminosity. Since the scattering fraction is typically less than 10 \\% for Seyfert 2 galaxies (\\cite{uen95}), the {\\LX}/{\\LOIII} should be less than 10 \\% of those of Seyfert 1 galaxies as is the case for NGC 1068 (\\cite{mul94}). However the observed X-ray to [OIII]$\\lambda$5007 luminosity ratio {\\LX}/{\\LOIII} is very similar to Seyfert 1 galaxies. Therefore the observed X-ray continuum is not likely to be due to a scattered component. Then the observed iron line should be emitted from the matter close to the nucleus in order to be ionized and/or broadened due to the Doppler effect. If the iron line is emitted by an accretion disk, a line profile with significant red tail is expected (\\cite{fab89}). On the other hand, the observed profile seems to be symmetric in shape although the statistics are limited. Broad lines with weaker red tails than Seyfert 1s are observed in AGNs with much higher luminosity; {\\LX}$>10^{44}$~{\\eps} (\\cite{na97b}). If the inner-most part of the accretion disk is almost fully ionized, the red component is expected to be very weak or absent. Thus the observed profile is consistent with the interpretation that the observed iron K emission is from an ionized disk. The obtained equivalent width ($\\sim 500$~eV for the Gaussian model) is rather large compared to that seen in most Seyfert 1 galaxies. If the disk is highly ionized, the fluorescence yield of iron increases and absorption by lighter elements decreases as light elements are almost completely ionized. In such a situation the equivalent width of an iron line can increase by a factor of two (Matt et al. 1993, \\cite{zyc94}). Therefore the large equivalent width is also naturally explained by an ionization effect. The ionization state of photoionized matter is determined by an ionization parameter $\\xi = L/nR^2$ (\\cite{kal82}), where $L$, $n$, and $R$ is the luminosity of ionizing photons, the number density of photoionized matter, and the distance from light source to photoionized matter, respectively. The X-ray luminosity of NGC 4579 is only $1.5\\times10^{41}$ {\\eps}, which is 1--3 orders of magnitude smaller than for Seyfert 1 galaxies, and the X-ray luminosity of M81, from which an iron line centered at $\\sim6.7$~keV is detected, is even lower ($\\sim 2\\times10^{40}$ {\\eps}). In order to photoionize iron atoms to be He-like, $\\xi$ should be at least $\\sim$ 500, while $\\xi<100$ is required for less ionized species ($<$ \\ion{Fe}{16}) which is probably appropriate for usual Seyfert 1 galaxies. Therefore $nR^2$ in the iron line emitting region should be more than two orders of magnitude smaller than that of luminous Seyfert 1 galaxies. An expected ionization parameter under an assumption of standard $\\alpha$ disk is calculated by \\cite{mat93}. According to their results, the ionization parameter has a strong dependence on the mass accretion rate $\\xi \\propto \\dot{m}^3$ (equations (5) and (6) in Matt et al. 1993), where $\\dot{m}$ is denoted in units of the critical accretion rate $\\dot{m} = L/L_{\\rm Edd}$. In order to ionize iron to He-like, $\\dot{m}$ should be at least 0.2 (Figs. 2 and 5 in Matt et al. 1993). However the order-of-magnitude estimate of the central black hole mass by \\cite{bar96} combined with the observed luminosity gives a significantly smaller value of $\\dot{m}$ $\\sim1\\times10^{-3}$. Then we cannot explain the very low luminosity and the ionized iron line at the same time in the standard disk model. This may suggest that the accretion processes in AGN is different in a very low luminosity situations with very small $\\dot{m}$. An advection dominated accretion flow (ADAF) model is proposed for AGNs specifically for objects radiating at very low Eddington ratio (e.g. $\\dot{m} \\sim 10^{-4}$ for NGC 4258, Lasota et al. 1996). In the model by Lasota et al. (1996), a standard disk is assumed outside of $r_{\\rm in}$ and an ADAF inside of $r_{\\rm in}$. In an ADAF, accreting matter is heated up to very high temperatures ($T_i\\sim10^{12}$K, $T_e\\sim10^{9}$K). However our detection of an iron line indicates the presence of highly ionized (but not fully ionized) matter surrounding a large solid angle viewed from the light source. This means that $r_{\\rm in}$ should be small and a geometrically thin disk is appropriate. Therefore the iron line in NGC~4579 cannot be explained solely by an ADAF model and the real situation in NGC~4579 may correspond to a condition near the transition from the $\\alpha$ disk to an ADAF. Future sophisticated modeling of accretion in LLAGNs and calculation of expected iron emission as well as precise measurements of an iron K line and mass determination by {\\HST} Space Telescope Imaging Spectrograph will be important to understand physical processes in extremely low luminosity AGNs.\\\\" }, "9804/gr-qc9804048_arXiv.txt": { "abstract": "We conduct a direct comparison of three different representative numerical codes for constructing models of rapidly rotating neutron stars in general relativity. Our aim is to evaluate the accuracy of the codes and to investigate how the accuracy is affected by the choice of interpolation, domain of integration and equation of state. In all three codes, the same physical parameters, equations of state and interpolation method are used. We construct 25 selected models for polytropic equations of state and 22 models with realistic neutron star matter equations of state. The three codes agree well with each other (typical agreement is better than 0.1 \\% to 0.01 \\%) for most models, except for the extreme assumption of uniform density stars. We conclude that the codes can be used for the construction of highly accurate initial data configurations for polytropes of index $N>0.5$ (which typically correspond to realistic neutron stars), when the domain of integration includes all space and for realistic equations with no phase transitions. With the exception of the uniform density case, the obtained values of physical parameters for the models considered in this paper can be regarded as ``standard'' and we display them in detail for all models. ", "introduction": "The physical state of the neutron star matter has not been fully understood yet because it is very difficult to investigate particle interactions beyond nuclear matter density ($\\varepsilon_{\\rm N}/c^2 \\sim 2 \\times 10^{14}$ g cm$^{-3}$) either from nuclear experiments or from nuclear theories, (here $\\varepsilon_{\\rm N}$ is the energy density of the nuclear matter and $c$ is the velocity of light). Given this situation, one promising approach to explore the behavior of very high density matter is to make use of the macroscopic quantities of neutron stars. In particular, the mass and the rotational period of neutron stars depend crucially on the softness of the equation of state (EOS) at very high densities (see e.g. Friedman et al. 1984, 1986, 1989), thus, observational constrains, matched with theoretical models, may help in reconstructing the equation of state of very high density matter. Given a particular equation of state, the mass of neutron stars varies with central energy density and always reaches a maximum. This implies that if the maximum mass of neutron star models constructed with a certain equation of state is smaller than the mass of observed neutron stars, that equation of state must be discarded. Currently, the largest accurately measured mass of slowly rotating neutron stars is $M_{\\rm BP} = 1.44 M_{\\odot}$, where $M_{\\rm BP}$ is the mass of one of the components of the binary pulsar PSR1913+16 (Taylor \\& Weisberg \\cite{tayl89}) and $M_{\\odot}$ is one solar mass. Individual masses of neutron stars have also been estimated in six other binary pulsars (Thorsett et al. \\cite{thor93}, Wolszczan \\cite{wol97}), as well as in six X-ray binaries (van Kerkwijk et al. \\cite{vankerk95}) but the accuracy is not as good as in PSR1913+16. Thus, equations of state which give larger masses than $M_{\\rm BP}$ for slowly rotating stars, can be valid as candidates for the real equation of state at very high densities. Since the maximum mass of neutron stars is smaller for more compressible (soft) equations of state than for less compressible (stiff) equations of state, the true equation of state at high densities cannot be extremely soft. On the other hand, stiff equations of state can be limited by considering the neutron star with the shortest rotational period, i.e. the most rapidly rotating pulsar. There exists a lower limit on the rotational period for each equation of state, because if the centrifugal force exceeds the self-gravity at the equatorial surface, no equilibrium states are allowed. The lower limit of the rotational period depends on the softness of the equation of state - the radius of neutron stars with softer equations of state is smaller, which allows for higher rotation rates. Thus, if very rapidly rotating neutron stars should be found, we could exclude most stiff equations of state. At the moment, the shortest period of observed pulsars is 1.56 ms, of PSR1937+21. Consequently, equations of state for which the shortest rotational period is larger than this value, must be excluded as candidates for the real equation of state for neutron star matter. The discussions above require us to make use of highly accurate schemes for constructing rotating neutron star models, in order to compute precise theoretical values of masses and rotational periods. Highly accurate relativistic equilibrium models are also needed as initial data for relativistic time-evolution codes (modeling of nonlinear pulsations, collapse and generation of gravitational waves). Recently, a number of groups have succeeded in constructing models of rapidly rotating neutron stars (Friedman et al. 1984, 1986, 1988, 1989, Eriguchi et al. \\cite{erig94}, Salgado et al. \\cite{salg94}, \\cite{salg94b}, Cook et al. \\cite{cook94b}, Stergioulas \\& Friedman \\cite{stag95} -- for a recent review see Stergioulas \\cite{S98}). However, the obtained models by those authors do not always agree with each other (see e.g. Friedman et al. \\cite{frie89}, Eriguchi et al. \\cite{erig94}, Salgado et al. \\cite{salg94}, Cook et al. \\cite{cook94b}, Stergioulas \\& Friedman \\cite{stag95}). Although Stergioulas \\& Friedman~(1995) have determined the cause of the discrepancy between models in Friedman et al.~(1989) and Eriguchi et al.~(1994), (which was due to the use of a slightly different equation of state table), the reasons of smaller differences which remain, even after using exactly the same equation of state, have not been clarified yet. This is because numerical techniques used in the different codes, such as the choice of parameters defining the model, the interpolation method, the method of integrating the field equations, a.s.o. are not the same. In this paper, three groups using their own codes (Komatsu et al. \\cite{koma89a}, Eriguchi et al. \\cite{erig94}, Salgado et al. \\cite{salg94}, Stergioulas \\& Friedman \\cite{stag95}) will decrease the differences between their results to a minimum possible, by tuning each code and using the same parameters, the same schemes of interpolation, the same equations of state, and so on. Since the basic schemes used by the three groups are different, it will be impossible to have exactly the same results and the relative differences between results are a measure of the accuracy of the codes. Models obtained with small relative differences between the three codes can be considered as ``standard\" models for each equation of state. Furthermore, this direct comparison allows us to investigate the effect that the choice of interpolation method, equation of state and domain of integration has on the accuracy of the codes. ", "conclusions": "\\subsection{Discussion} \\subsubsection{Metric Potentials} As redshift factors differ by about 10 \\% for the constant density, relativistic model N00$rr$ between the three codes (while the agreement global quantities is within a few \\%) we compare directly the local values of metric potentials for several models. Figures 1 to 16 show the four metric potentials (upper panel) and the relative differences in them between BGSM and KEH(OR) (middle panel) and KEH(SF) and BGSM (lower panel) for the models N05$mr$, N15$mr$, L(L)$mr$ and WFF(FPS)$br$. The metric potentials are graphed against the coordinate $r$ in the equatorial plane ($\\theta= \\pi/2$, solid line) and along the axis of rotation ($\\theta=0$, dashed line). The range of the coordinate $r$ is the twice the equatorial radius of the star. In general, the agreement in the local values of the metric potentials reflects the agreement in the computed physical parameters of models. In these graphs, several significant behaviors can be pointed out: First, there are high frequency and small amplitude oscillations at the inner part of the stars for all models. Second, the differences between the results of KEH(OR) and those of the other two codes are growing outside the stars as $r$ increases. Third, although the differences between the KEH(SF) and BGSM codes are very small for models N05$mr$, N15$mr$ and WFF(FPS)$br$, there appear larger differences for the stiff model L(L)$mr$. Fourth, there appears a larger amplitude oscillation in the metric potential $\\omega$ on the axis of rotation, close to the surface. The first behavior is due to the integration scheme of the KEH code, i.e. the Simpson scheme. In general, the Simpson scheme gives results with higher precision, compared with those obtained by the trapezoidal scheme. However, in the KEH scheme, the integrands contain nonsmooth functions with respect to the radial coordinate, because of the nature of the Green's functions. The non-uniform distribution of the weight factor in Simpson's scheme for nonsmooth functions results in oscillating behaviors with very small amplitudes, which cannot be noticed in the behavior the integrated quantities. The second behavior in the original KEH code is caused by the \"truncation\" of the domain of integration at a finite distance from the star, instead of integrating over the whole space. The large differences in the metric potentials between KEH(SF) and BGSM for EOS L, could be accounted to the stiffness of the equation of state, but the differences between KEH(OR) and BGSM for the same model are not as large, and we have not an explanation for that. The oscillations in $\\omega$ on the axis of rotation near the surface are present also for the soft $N=1.5$ polytropes, while for $N=0.5$ they are larger. This indicates that terms in the field equations for $\\omega$ are very sensitive to the presence of the surface and the accompanying Gibbs phenomenon. Even for $N=1.5$ polytropes, where the density goes to zero smoothly at the surface, there is a small scale Gibbs phenomenon, due to the finite number of grid points used to represent the region of integration. \\subsubsection{Method of Interpolation} An important factor for the local accuracy of models constructed with realistic equations of state is the method of interpolation of the energy vs. pressure data given in an EOS table. While global quantities are not affected significantly, the virial identities for realistic EOSs, are sensitive to the interpolation scheme This can be considered to reflect the nature of the interpolation scheme as mentioned before. If we define the enthalpy ($H$) by \\begin{equation} H \\equiv \\ln \\left( { \\varepsilon + p \\over \\rho c^2} \\right), \\end{equation} the Gibbs-Duhem relation, which follows directly from the first law of thermodynamics, implies \\begin{equation} {dp\\over dH} = \\varepsilon + p \\ . \\end{equation} In the cubic Hermite interpolation, the Gibbs-Duhem relation is used to replace by $\\nabla H$ the term $\\nabla p /(\\varepsilon+p)$ which appears in the hydrostationary equilibrium equation. If the tabulated function $p(H)$ fails to satisfy the above relation, then the hydrostationary equilibrium equation, which is derived from the Bianchi identity, is only approximately verified by the numerical solution, which results in increased error in the GRV2 and GRV3 virial identities. The four point Lagrange interpolation does not satisfy the Gibbs-Duhem relation because it only reproduces the values of the discrete points, but there is no guarantee for the reproduction of the derivatives. This explains why the GRV2 and GRV3 errors are bad, even in the nonrotating case (GRV2 = 3E-03, GRV3 = 1E-02 for model L$sr$) as compared to ${\\rm GRV2}\\sim 10^{-14}$ for polytropic models (see e.g. Bonazzola et al. \\cite{bona93}). The GRV2 and GRV3 error indicators thus do not reflect the precision of the code but the bad thermodynamical behavior of the tabulated EOS. The advantage of the cubic Hermite interpolation is that the Gibbs-Duhem relation is automatically satisfied because this interpolation reproduces not only the values themselves but also the derivatives (Swesty 1996). Moreover, in our case, the energy density and the baryon number density are obtained by \\begin{eqnarray} \\varepsilon & = & {p \\over H} {d\\log p\\over d\\log H} - p, \\\\ n & = & {\\varepsilon + p \\over m_{\\rm B} c^2} \\exp(-H) \\ . \\end{eqnarray} Because of these equations, the Gibbs-Duhem relation is satisfied in every point. Note also that the value of $\\varepsilon$ obtained in this way coincides exactly with $\\varepsilon_i$ at the points in the tabulated equation of state. \\subsection{Conclusion} The comparison of three different codes for constructing rapidly rotating relativistic neutron star models demonstrates that the BGSM and KEH schemes used are highly accurate for typical polytropic models - when the field equations are solved to infinity - and for models constructed with realistic equations of state, that do not have phase transitions. If one approximates neutron stars as constant density stars, then Gibbs phenomena at the discontinuous surface reduce the accuracy of the computed models. If high accuracy in such models and in models with phase transitions is desired, then modified numerical schemes - free of Gibbs phenomena - need to be used. Such numerical schemes could employ, for example, surface fitted coordinates. Such a scheme has been presented recently by Bonazzola et al. (1998a) in the framework of spectral methods and looks promising for rotating stellar models. Further, we demonstrated that the metric potentials are subject to various local oscillatory behaviors, even if integrated quantities have very good accuracy. This observation is important for the effort of constructing initial data for the numerical evolution of rotating relativistic neutron star models. \\begin{figure} \\resizebox{\\hsize}{10.5cm}{\\includegraphics{zeta_19.eps}} \\caption{Same as Figure 1 but for the metric potential $\\zeta$ of model N15$mr$.} \\end{figure} \\begin{figure} \\resizebox{\\hsize}{10.5cm}{\\includegraphics{zeta_27.eps}} \\caption{Same as Figure 1 but for the metric potential $\\zeta$ of model WFF(FPS)$br$.} \\end{figure} \\begin{figure} \\resizebox{\\hsize}{10.5cm}{\\includegraphics{zeta_4.eps}} \\caption{Same as Figure 1 but for the metric potential $\\zeta$ of model N05$mr$.} \\end{figure} \\begin{figure} \\resizebox{\\hsize}{10.5cm}{\\includegraphics{zeta_26.eps}} \\caption{Same as Figure 1 but for the metric potential $\\zeta$ of model L(L)$mr$.} \\end{figure}" }, "9804/gr-qc9804081_arXiv.txt": { "abstract": "In this paper we investigate extended inflation with an exponential potential $V(\\sigma)= V_0\\,e^{-\\kappa\\sigma}$, which provides a simple cosmological scenario where the distribution of the constants of Nature is mostly determined by $\\kappa$. In particular, we show that this theory predicts a uniform distribution for the Planck mass at the end of inflation, for the entire ensemble of universes that undergo stochastic inflation. Eternal inflation takes place in this scenario for a broad family of initial conditions, all of which lead up to the same value of the Planck mass at the end of inflation. The predicted value of the Planck mass is consistent with the observed value within a comfortable range of values of the parameters involved. ", "introduction": "The extended inflation action is given by \\cite{extended} \\begin{equation} \\label{a} S =\\int d^{4}x \\,\\sqrt{-g}\\left[\\Phi R - {\\omega\\over\\Phi}(\\partial \\Phi)^{2} - \\frac{1}{2}(\\partial \\sigma)^{2} \\\\ - V(\\sigma)\\right] , \\end{equation} where $R$ is the curvature scalar and the potential is $V(\\sigma) = V_0 e^{-\\kappa\\sigma}$. The coupling $\\omega$ plays a similar r\\^{o}le as that of the coupling functions $B_i(\\Psi)$ of the dilaton field $\\Psi$ in string theory. Based on this analogy, several authors have investigated the so-called {\\it hyperextended inflation} models \\cite{extended2,extended3,mikel2}, where $\\omega$ becomes dependent on the BD field. In this paper however, we will merely examine the $\\omega={\\rm const}$ model. The BoI boundary is given by $V(\\sigma)= M^{4}_{\\rm p}(\\Phi)$ or equivalently, \\begin{equation} \\label{c1} \\Phi = \\frac{V_0^{1/2}}{16\\pi}\\,e^{-\\kappa\\sigma/2}. \\end{equation} The BoI is the quantum limit where the metric fluctuations become significant and the inflaton field cannot take the values for which the potential is above this boundary. The EoI boundary is on the other hand \\begin{equation} \\label{d} \\frac{1}{2}\\dot \\sigma^{2} + \\omega\\frac{\\dot \\Phi^{2}}{\\Phi} \\approx V(\\sigma). \\end{equation} The equations of motion in a flat FRW background are \\begin{equation} \\label{e1} \\left(D^2+\\frac{1}{2\\omega}R\\right)\\,\\Phi=0\\,, \\end{equation} \\begin{equation} \\label{e2} D^{2}\\sigma = -V^{\\prime}(\\sigma)\\,, \\end{equation} \\begin{equation} \\label{e3} H^{2}+H\\frac{\\dot \\Phi}{\\Phi} = \\frac{\\omega}{6} \\left({\\dot\\Phi\\over\\Phi}\\right)^2 + \\frac{1}{6\\Phi} \\left[\\frac{1}{2}\\dot\\sigma^{2} + V(\\sigma)\\right] , \\end{equation} and the differential operator $D$ is defined \\begin{equation} \\label{f} D^{2} \\equiv \\partial^{2}_{t} + 3H\\partial_{t} . \\end{equation} In the slow-roll approximation, $\\ddot \\Phi \\ll H\\dot\\Phi \\ll H^{2}\\Phi$ and $\\dot \\sigma^{2}+2\\omega\\, \\dot\\Phi^{2}/\\Phi \\ll 2V(\\sigma)$, (\\ref{e1})-(\\ref{e3}) read \\begin{eqnarray} \\label{g1} \\frac{\\dot\\Phi}{\\Phi} &=&2\\frac{H}{\\omega} \\,, \\\\ \\dot \\sigma &=& -\\frac{1}{3H}V^{\\prime}(\\sigma) \\,,\\\\ H^{2} &=&\\frac{1}{6\\Phi}V , \\end{eqnarray} and the curvature scalar is given by $R=-12H^{2}$. The slow-roll equations (\\ref{g1}) enable us to rewrite (\\ref{d}): \\begin{equation} \\label{h} \\Phi_* =\\left(3-\\frac{2}{\\omega}\\right)\\frac{1}{\\kappa^{2}}, \\end{equation} where the $*$ subindex denotes the value at the end of inflation. Hence, $\\Phi_*$ is independent of $\\sigma$ and, for reasonably large $\\omega$, it is solely determined by the slope of the potential. The condition $\\Phi>0$ also imposes the constraint $\\omega$, as can be seen in Fig.~1, such that the range $0<\\omega<2/3$ is excluded to prevent imaginary values of the Planck mass. The classical trajectories of the fields are given by the following conservation law \\cite{mikel2} \\begin{equation} \\label{h1} \\frac{d}{dt}\\left[\\omega\\Phi + \\int d\\sigma \\frac{V(\\sigma)}{V^{\\prime}(\\sigma)} \\right]=0, \\end{equation} which in the case of the exponential potential yields \\begin{equation} \\label{h2} \\Phi= \\frac{\\sigma}{\\kappa \\omega} + \\left(\\Phi_0- \\frac{\\sigma_0}{\\kappa \\omega}\\right). \\end{equation} In Fig.~2 we have put together the BoI and EoI curves and the classical trajectories of the fields on the ($\\sigma$,$\\Phi$) plane, i.e. (\\ref{c1}),(\\ref{h}) and (\\ref{h2}) respectively. It can be seen in the figure that inflation takes place in the region enclosed by BoI and EoI to the right of the intersection point $A$. The trajectories given by (\\ref{h2}) are straight lines parallel to the segment $BC$, and the fields drift along these curves in the direction $B\\to C$ during the course of inflation. The region enclosed by BoI and EoI to the left of $A$ does not undergo inflation, because the orientation of the classical trajectories is such that the fields would move from EoI towards BoI, which is not an acceptable solution. In addition to the classical trajectories quantum diffusion is responsible for the jumps of the fields between neighbouring classical trajectories. It can be seen that, unlike with powerlaw potentials, for which $\\sigma$ decreases as $\\Phi$ increases during the course of inflation, in the case of the exponential potential both fields increase during the slow roll. \\begin{figure}[t] \\centering \\leavevmode\\epsfysize=5.5cm \\epsfbox{omega.ps}\\\\ \\vskip 0.2cm \\caption[fig1]{BD field at EoI, $\\Phi_*$, vs. $\\omega$ for an arbitrary value of $\\kappa$. $\\Phi_*$ is given in units of $\\kappa^{-2}$. Inflation takes place in the range $\\omega<0$ or $\\omega\\geq 2/3$.} \\end{figure} \\begin{figure}[t] \\centering \\leavevmode\\epsfysize=5.5cm \\epsfbox{fields.ps}\\\\ \\vskip 0.2cm \\caption[fig2]{Predicted BoI and EoI curves, dashed and thick solid curves respectively. Classical trajectories are straight lines parallel to $BC$. At the intersection point $A$ the onset and end of inflation coincide. $\\sigma_{\\rm max}$ determines the scale of validity of the slow-roll approximation. Inflation takes place within the region enclosed by BoI, EoI and $\\sigma\\approx\\sigma_{\\rm max}$.} \\end{figure} The EoI boundary (\\ref{h}) gives a definite and unique prediction for $\\Phi_*$, and also it implies that if $\\Phi_0>\\Phi_*$ inflation will not occur. In the case of $0<\\Phi_0\\lsim\\Phi_*$, inflation takes place for values of the inflaton \\begin{equation} \\label{sigma0} \\sigma_0 \\gsim -\\frac{2}{\\kappa}\\, {\\rm log}\\left({16\\pi\\Phi_0 \\over V_0^{1/2}}\\right). \\end{equation} Naturally if $\\Phi_0=\\Phi_*$, then the RHS of (\\ref{sigma0}) is $\\sigma_A$, the value of the field at the intersection point $A$ of BoI and EoI in Fig.~2. It must be noted that the slow-roll approximation does not hold in the case of the exponential potential for arbitrarily large values of $\\sigma$. For a given value of $\\kappa$ it is straightforward to compute $\\sigma_{\\rm max}$ for which the potential and kinetic energy of the fields are comparable and thus the slow-roll conditions break down. It is easy to show from (\\ref{d}) that this scale is \\begin{equation} \\label{smax} \\sigma_{\\rm max}\\approx \\left({3\\omega-2\\over\\kappa}\\right) . \\end{equation} Therefore it follows that the EoI boundary does not span from $\\sigma_A$ to infinity, but inflation will occur within a finite region $\\sigma_A\\lsim\\sigma\\lsim\\sigma_{\\rm max}$. ", "conclusions": "In this paper we have examined extended inflation with an exponential potential. The remarkable feature of this model is the prediction of a constant distribution of the Planck mass at the end of inflation, given by (\\ref{h}). The parameter $\\kappa$ of the theory is therefore estimated via the observed Planck mass in this region of the universe, which in turn fixes the parameter $\\sigma_{\\rm max}$ that determines the range of values of $\\sigma$ for which inflation takes place. The amplitude of the potential $V_0$ is left unconstrained by astrophysical bounds on the spectrum of fluctuations, as described by the argument given in \\S 4. The dynamics as is given in \\S 2 and the likelihood distributions in \\S 3 are shown to be insensitive to the numerical value of this parameter. As is shown in Fig.~2, the BoI and EoI curves in this model cross at $\\sigma=\\sigma_A$ and the area enclosed between them is thus infinite. However the breakdown of the slow-roll approximation for the exponential potential over the range $\\sigma\\gsim \\sigma_{\\rm max}$ (where $\\sigma_{\\rm max}$ is given by (\\ref{smax})) implies that in practical terms only a finite region of the ($\\sigma$,$\\Phi$) plane undergoes inflation. In the classification of \\cite{V} this means that the exponential potential is {\\it class I}, i.e. the values of the fields at EoI remain finite. \\subsection*" }, "9804/astro-ph9804239_arXiv.txt": { "abstract": "This study is the first known attempt to search for gamma-ray burst repeaters combining data from gamma-ray experiments flying on board different satellites and making use of information derived from the bursts detected simultaneously by all the experiments. The proposed method is suitable to correlate GRB data provided by experiments that overlap partially or totally in time. As an application of this method we have correlated the positions of 57 gamma-ray bursts observed by WATCH/GRA\\-NAT and WATCH/EURECA with 1905 bursts detected by BATSE. Comparing the so-called~~~``added correlation''~~~between~~~the WATCH and BATSE bursts with that obtained with simulated WATCH catalogues, we conclude that there is no indication of recurrent activity of WATCH bursts in the BATSE sample. We derive an upper limit of $15.8\\%$, with a confidence level of $94\\%$, for the number of WATCH gamma-ray bursts that could represent a population of repeaters in the BATSE sample. ", "introduction": "Despite the advances carried out so far, the origin of the gamma-ray bursts (hereafter GRBs) remains unknown. The identification of absorption lines in the optical spectrum of GRB 970508 strongly supports models arising from sources at cosmological distances (Metzger et al. 1997), but there is still a lack of knowledge on the mechanisms originating these enigmatic phenomena. One of the most important clues that could clarify the nature of the GRBs would be the detection of a repeater behaviour. Initial studies showed an apparent evidence of repetition for the BATSE 1B catalogue (Quashnock and Lamb 1993), suggesting that it would be possible to have an excess of pairs of GRBs clustered in both time and space (Wang and Lingenfelter 1995). This fact was not confirmed by the work carried out using the BATSE 2B catalogue (Brainerd et al. 1995), although other studies provided marginal evidence for both temporal and angular clustering (Petrosian and Efron 1995). Analyses based on autocorrelations with data from the BATSE 3B catalogue did not find any evidence of repetition (Bennett and Rhie 1996) and have imposed several constraints to the number of repeaters (Tegmark et al. 1996). Finally, recent studies confirm the lack of repetition in the 4B catalogue and lead to an upper limit to the repetition rate of $ 0.04$ burst source$^{-1}$ yr$^{-1}$ (Hakkila et al. 1997). The BATSE~4B catalogue was obtained by the BATSE experiment on board the {\\it CGRO} satellite and contains 1637 GRBs detected from April 1991 to August 1996 (Paciesas et al. 1998). The BATSE experiment consists of eight identical detector modules, placed at the corners of the {\\it CGRO} spacecraft and covering energy channels from $\\sim 25$ keV to $\\sim$ 2 MeV. It provides error boxes with a minimum radius of $1.6^{\\circ}$ (1$\\sigma$ confidence level, Fishman et al. 1994). BATSE is detecting bursts at a rate of 0.8 bursts per day. The bursts are daily added to the so-called Current GRB Catalogue, which contains the BATSE~4B catalogue plus all bursts detected after August 1996. When this study was started, the catalogue contained 1905 sources; this sample constitutes the basis of the present work. The WATCH X-ray all-sky monitor is based on the rotation modulation principle (Lund 1986). The instrument has a circular field of view of 4 steradians and an effective area of $\\sim$ 30 cm$^2$ (averaged over the field of view). Position sensitivity is achieved using the rotation collimator principle, with the collimator grids rotating with a frequency $\\omega$=1 Hz. The phoswich detectors consist of interleaved scintillator-strips of NaI and CsI crystals. The geometric area of the scintillator is 95~cm$^2$. Four units were mounted on board the Soviet {\\it GRANAT} satellite in a tetrahedral configuration covering the whole sky, and one unit on board the European Space Agency {\\it EURECA} spacecraft. The total energy range is 8-80~keV, therefore overlapping with the lower BATSE energy band. WATCH\\-/GRANAT detected bursts in 1990-94 and WATCH/EURECA in 1992-93, thus both experiments also overlapped in time with BATSE. One of the main advantages of WATCH was the capability of locating bursts with relatively small error boxes ($3\\sigma$ error radii with $\\sim$ 1$^{\\circ}$) (Brandt et al. 1990). WATCH\\-/GRANAT detected 47 GRBs in this period and WATCH\\-/EURECA 12 (Castro-Tirado et al. 1994, Brandt et al. 1994, Sazonov et al. 1998). Two GRBs (GRB 920814 and GRB 921022) were detected by both the WATCH\\-/GRANAT and WATCH\\-/EURECA experiments. Therefore, the sample of WATCH GRBs used in this study comprises 57 GRBs: 45 WATCH\\-/GRANAT bursts, 10 WATCH\\-/EURECA bursts and the above-mentioned two GRBs. BATSE also detected 27 of them. Fig.~\\ref{figure1} shows the sample of 57 WATCH GRBs used in this study. \\begin{figure*} \\centering \\resizebox{!}{!}{\\includegraphics[width=\\hsize,totalheight=10.7cm]{7502.f1}} \\caption{Error boxes for the 57 GRBs detected by WATCH, represented in galactic coordinates. The sample contains 45 GRBs detected by WATCH/GRANAT, 10 by WATCH/EURECA and two localized by both experiments at the same time. The typical radii of the error boxes are $\\sim 1 ^{\\circ}$, with a $3 \\sigma$ confidence level.} \\label{figure1} \\end{figure*} The distribution of time amplitudes for GRBs shows two classes of bursts: a) durations shorter than $\\sim$ 2 s and b) longer than $\\sim$ 2 s (Kouveliotou et al. 1993). It was noticed that the energy spectra of the short bursts were generally harder than those of the long ones (Kouveliotou et al. 1993, Lestrade et al. 1993). The fraction of short events in the WATCH sample is smaller than that in the 4B catalogue. This fact can be justified by at least three selection effects: i) The availability of WATCH for localizing sources is governed by the rotation speed of the collimator grids (1~Hz). So, a source needs to be bright enough for at least one rotation of the modulation collimator in order to be localized, implying a burst duration longer than 1~s. In contrast, the BATSE experiment is able to detect bursts with durations as short as 64 ms. ii) The low energy band of the WATCH experiment ($\\sim$8-20~KeV) is sensitive to the soft GRBs, below the BATSE lower limit ($\\sim$25 KeV), which generally belong to the class of bursts with durations longer than 2 s. iii) On the other hand, since WATCH is about an order of magnitude less sensitive than the large-area detectors of BATSE, the WATCH catalogue contains bursts which are brighter than those in the BATSE sample. The above three reasons~~explain~~why~~the GRBs~~in the WATCH sample are longer, softer and brighter than the average BATSE 4B bursts. This study is the first known attempt to search for repeaters combining data of $\\gamma$-ray experiments flying on board different satellites. The method proposed makes use of the so-called ``simultaneous bursts'' and is suitable to correlate GRB data provided by experiments that overlap partially or totally in time. In the future, this work could also be used to detect systematic pointing errors between different $\\gamma$-ray experiments, allowing to improve their capability for locating GRBs. ", "conclusions": "In this study we have developed a method that allows us to search for GRBs common to two catalogues of sources, each one based on a different instrument. The method makes use of the GRBs detected simultaneously by both experiments, so it is necessary that the experiments overlap in time. We have applied~~the~~method~~to the WATCH (WATCH/GRANAT + WATCH\\-/EU\\-RE\\-CA) and BATSE (BATSE 4B + bursts detected after August 1996) catalogues. We conclude that there is no evidence of recurrent activity of WATCH bursts in the BATSE sample. We claim (with a $94\\%$ confidence level) that no more than a $15.8\\%$ of the 57 GRBs detected by WATCH are present in the sample of 1905 BATSE bursts (excluding the simultaneous bursts). However, the possibility of finding repeaters in each single catalogue cannot be ruled out. Our results support models which do not predict repetitions of GRBs (for instance the merging of neutron stars at cosmological distances)." }, "9804/astro-ph9804180_arXiv.txt": { "abstract": "s{ We address the problem of map-making with data from the \\ps\\ High Frequency Instrument, with an emphasis on the understanding and modelling of instrumental effects, and in particular that of sidelobe straylight. } ", "introduction": " ", "conclusions": "We have shown that the scan strategy of \\ps\\ along rings on the sky allows to decompose the problem of converting data streams into CMB anisotropy maps in two independent steps. This makes the problem tractable numerically, and helps in analysing and monitoring the impact of systematic effects. In particular, a promising method for the identification and removal of sidelobe signals has been developped." }, "9804/astro-ph9804149_arXiv.txt": { "abstract": "We review recent results of Sunyaev-Zel'dovich effect (SZE) observations toward galaxy clusters. Using cm-wave receivers mounted on the OVRO and BIMA mm-wave arrays we have obtained high signal to noise images of the effect for more than 20 clusters. We present current estimates of the Hubble constant and cosmological parameters and discuss the potential of conducting statistical studies with large SZE cluster samples. ", "introduction": "Over the last few years there has been a tremendous increase in the study of galaxy clusters as cosmological probes, initially through the use of X-ray emission observations, and in recent years, through the use of Sunyaev-Zel'dovich effect (SZE). Briefly, the SZE is a distortion of the cosmic microwave background (CMB) radiation by inverse-Compton scattering of thermal electrons within the hot intercluster medium (Sunyaev \\& Zel'dovich 1980, see Birkinshaw 1998 for a recent review). The change in the CMB brightness temperature observed is: \\begin{equation} \\frac{\\Delta T}{T_{\\rm CMB}} = \\left[ \\frac{x (e^{x}+1)}{e^{x}-1} -4 \\right] \\int \\left(\\frac{k_B T_e}{m_e c^2}\\right) n_e \\sigma_T dl, \\end{equation} where $x = h \\nu/k_B T_{\\rm CMB}$, and $n_e$, $T_e$ and $\\sigma_T$ are the electron density, electron temperature and the cross section for Thomson scattering. The integral is performed along the line of sight through the cluster. The other important observable of the hot intercluster gas is the thermal Bremsstrahlung X-ray emission, whose surface brightness $S_X$ can be written as: \\begin{equation} S_X = \\frac{1}{4 \\pi (1+z)^3} \\int n ^{2}_{e} \\Lambda_e dl, \\end{equation} where $z$ is the redshift and $\\Lambda_e(\\Delta E,T_e)$ is the X-ray spectral emissivity of the cluster gas due to thermal Bremsstrahlung within a certain energy band $\\Delta E$. By combining the intensity of the SZE and the X-ray emission observations, and knowing the cluster gas temperature $T_e$, the angular diameter distance to the cluster can be derived due to the different dependence of the X-ray emission and SZE on the electron density, $n_e$. Combining such distance measurements with redshift allows a determination of the Hubble constant, H$_0$, as a function of certain cosmological parameters (e.g., Hughes \\& Birkinshaw 1998a). If distance measurements for a sample of clusters exist, then the angular diameter distance with redshift relation can be used to put constraints on the cosmological models, similar to current supernovae constraints at high redshift. ", "conclusions": "" }, "9804/astro-ph9804133_arXiv.txt": { "abstract": "A new method of shower-image analysis is presented which appears very powerful as applied to those Cherenkov Imaging Telescopes with very high definition imaging capability. It provides hadron rejection on the basis of a single cut on the image shape, and simultaneously determines the energy of the electromagnetic shower and the position of the shower axis with respect to the detector. The source location is also reconstructed for each individual $\\gamma$-ray shower, even with one single telescope, so for a point source the hadron rejection can be further improved. As an example, this new method is applied to data from the C{\\small AT} (Cherenkov Array at Th\\'emis) imaging telescope, which has been operational since Autumn, 1996. ", "introduction": "Many of the sources in the E{\\small GRET} catalogue \\cite{egret} have well-identified radio, optical, or X-ray counterparts, which give the source position to an accuracy much better than can be achieved by {\\small ACT} telescopes. Such sources with known position are usually placed at the centre of the field of an imaging telescope. However, many unidentified sources in the E{\\small GRET} catalogue are located in error boxes with typical size of 1$^\\circ$. Ground-based Cherenkov detectors are, in principle, able to localize such sources with a higher precision. In order to observe these sources, different methods have been developed by imaging Cherenkov telescope groups. Stereoscopy is the most direct way to find the direction of a source, but requires at least two telescopes and reduces the collection area \\cite{hegratel}. For single-telescope experiments with sufficient background rejection on an image-shape criterion, it is possible to perform de-localized analyses assuming the source at the different points of a grid covering the field of view or by examining the distribution of the intersections of the image axes \\cite{grid}. However, an event-by-event analysis method such as that described here is preferable since in the former methods the signal is more easily drowned-out by the background. The method has been tested on simulated $\\gamma$-images provided by the Monte-Carlo simulation program described above and a realistic simulation of the detector response, including the measured variation in collection efficiency and gain between the phototubes, and measured wavelength response of the mirrors and Winston cones. For optimization of the cut values, these simulated $\\gamma$-images have been used together with the real background events from data from off-source runs with the C{\\small AT} imaging telescope. Gammas from a point-like source with the Crab nebula spectrum \\cite{catcrab} (see equation (1) above) were simulated at various elevations. The capability of the method both for source detection and for source spectrum measurement have been examined. \\subsection{Source detection} As applied to the data, the method consists of minimizing the $\\chi^2$ with respect to $E_\\gamma$, $D$, $\\vec\\xi$, and $\\phi$. Fig~\\ref{fig:chi2} shows the $\\chi^2$ probability distributions obtained with this fit for the simulated $\\gamma$-ray events and real background events. \\begin{figure} \\epsfxsize=14.5 cm \\leavevmode \\centering \\epsffile[0 25 590 520]{fig6.eps} \\caption{$\\chi^2$ probability distribution for a fit with the source coordinates considered as free parameters (constraint is from shape alone): a) vertical $\\gamma$-ray showers; b) real background (off-source data) showers. The upper line in each figure is for all events above threshold, the shaded distributions for events with a fitted energy, $E_{\\mathrm{f}}$, greater than $350\\:{\\mathrm {GeV}}$ and a fitted impact parameter, $D_{\\mathrm{f}}$, between 30 and $125\\:{\\mathrm{m}}$.} \\label{fig:chi2} \\end{figure} A cut on the $\\chi^2$ probability value, $P(\\chi^2)$, provides a selection of $\\gamma$-like events on the basis of the image shape alone. The reconstructed angular origins obtained for simulated $\\gamma$-events accumulate around the actual source position which in this case is at the centre of the field. The dispersion around the actual source position has a typical {\\small RMS} spread of $0.14^\\circ$. For each event, the accuracy of the angular origin determination is better by a factor two in the direction perpendicular to the image axis than in the direction of the image axis (Fig.~\\ref{fig:pointerr}). \\begin{figure} \\epsfxsize=14.5 cm \\leavevmode \\centering \\epsffile[20 25 540 230]{fig9.eps} \\caption{Distributions of longitudinal and transverse errors (with respect to the image major axis) in the reconstruction of the source position for vertical showers with $P(\\chi^2)>0.2$. The small bump at negative values of longitudinal error results from wrong direction reconstruction (mainly for events close to the energy threshold). Shaded histograms correspond to events with a fitted energy greater than $350\\:{\\mathrm GeV}$.} \\label{fig:pointerr} \\end{figure} The {\\small RMS} longitudinal error typically varies from $0.2^\\circ$ to $0.1^\\circ$ as the energy varies from the threshold to $2\\:{\\mathrm {TeV}}$. The angular origins obtained for background events are spread over the whole field with an approximately Gaussian distribution with a $1.8^\\circ$ {\\small FWHM}. Since this distribution is fairly flat, 2-dimensional skymaps of the angular origins of the showers could be used for source detection, as can be seen from the reconstructed positions in data taken on Markarian 501 in Fig.~\\ref{fig:m501}. \\begin{figure} \\epsfxsize=14.5 cm \\leavevmode \\centering \\epsffile[0 25 580 520]{fig10.eps} \\caption{The distribution of reconstructed angular origins for the data from a 30-minute run on Markarian 501 from April 16, 1997, for events with $P(\\chi^2)>0.2$. The concentric circles represent the trigger region and the small-pixel region, respectively. During the run, the optic axis described the small arc indicated due to the mechanical flexibility of the structure, which is monitored as described in [9]. The number of events reconstructed in each bin of $(0.05{^0})^2$ is shown. No background subtraction has been performed.} \\label{fig:m501} \\end{figure} The errors in angular reconstruction given above are for a single shower; a point source with poorly defined position could thus be localized to $\\sim 1-2'$ with the combination of $\\sim 100$ such events. For the present, a conservative procedure of monitoring the background is used, based on the pointing angle $\\alpha$, similar to the ``orientation'' of Whipple \\cite{scuts}: $\\alpha$ is the angle at the image barycentre between the actual source position and the reconstructed source position. The pointing angle does not use the full information contained in the results of the fit, but has a fairly flat distribution from $0^\\circ$ to about $120^\\circ$ for background events, which allows the background level to be easily monitored. The cut on $\\alpha$ is more efficient than a cut on the angular distance between the source position and the reconstructed $\\gamma$ origin since, as seen in Fig.~\\ref{fig:pointerr}, the position reconstructed is not symmetric about the source position. The distribution of $\\alpha$ for $\\gamma$-events from a Crab-like source exhibits a peak at $0^\\circ$ (Fig.~\\ref{fig:alpha}) and a small accumulation at $180^\\circ$ corresponding to events which are wrongly found to point away from the source. \\begin{figure} \\epsfxsize=14.5 cm \\leavevmode \\centering \\epsffile[5 25 580 520]{fig7.eps} \\caption{Distribution of the pointing angle $\\alpha$ for events with $P(\\chi^2)>0.2$ (constraint from shape alone): a) vertical $\\gamma$-ray showers; b) Real background (off-source data) showers. The upper line is for all events above threshold, the shaded histograms for events with a fitted energy greater than $350\\:{\\mathrm {GeV}}$.} \\label{fig:alpha} \\end{figure} Around 17\\% of the $\\gamma$-events from a Crab-like source are in this situation. The proportion of events with a wrongly reconstructed direction decreases with increasing energy, from 22\\% at $200\\:{\\mathrm {GeV}}$ to 9\\% at $600\\:{\\mathrm {GeV}}$ and 4\\% at $1\\:{\\mathrm {TeV}}$. The significance of a signal is calculated using the usual formula: $(ON-OFF)/\\sqrt{ON+OFF}$ \\cite{lima}, assuming equal time on and off-source. The significance per hour on a simulated Crab-like source at zenith has been calculated for various cut values on $\\alpha$ and $P(\\chi^2)$ (Fig.~\\ref{fig:signif1}.a). \\begin{figure}[t] $$ \\epsfxsize=6.5 cm (a)\\epsffile[40 30 540 490]{fig8.eps} \\; \\epsfxsize=6.5 cm (b)\\epsffile[40 30 540 490]{fig11.eps} $$ \\vspace{-1cm} \\caption{Detection significance for a Crab-like source at zenith in one hour of observation, shown as a function of the two cuts on $P(\\chi^2)$ and $\\alpha$: (a) for a source at the centre of the camera; (b) for a source at $1^\\circ$ from the centre. The full lines indicate contours of equal significance; dotted lines show fixed $\\gamma$-ray selection efficiency.} \\label{fig:signif1} \\end{figure} The best result in terms of both significance and efficiency for $\\gamma$-events is obtained for a $P(\\chi^2)>0.2$ and $\\alpha<6^\\circ$, which gives $5.1\\sqrt{t}\\ \\sigma$ where $t$ is the on-source observation time in hours, retaining 34\\% of the $\\gamma$-events while giving a rejection factor of 120 on background events. This rejection factor is smaller than for some comparable experiments as there is a large rejection factor at the trigger level, allowing a moderate background rate of $15\\:{\\mathrm {Hz}}$ at the zenith. At $45^\\circ$ from zenith the best significance for the same cuts falls to $2.7\\sqrt{t}\\ \\sigma$. This is essentially due to the higher energy threshold, leading to a lower event rate; on the other hand, for the same $\\gamma$-ray selection efficiency the background rejection factor is comparable to that at zenith. In order to estimate the efficiency of the $\\chi^2$-method in the case of a source with a poorly-defined position, a simulated Crab-like source has been set on the edge of the trigger area ($1^\\circ$ from the centre). In this case, the equivalent detection area is divided by a factor of the order of two. Even for a source with known position, when the source is not at the centre of the field it is possible to use one side of the camera as the off region for the other side \\cite{offc}. For a Crab-like source at zenith, the same cuts in $P(\\chi^2)$ and $\\alpha$ as for a source in the centre of the camera give a $2.3\\sqrt{t}\\ \\sigma$ significance, with a selection efficiency for $\\gamma$'s of 36\\% and a rejection factor of 116 above threshold (Fig.~\\ref{fig:signif1}.b). This means that the on-source run time has to be five times larger for a source on the edge of the trigger area than for a source at the centre of the field for the same significance. The corresponding significance obtained at $45^\\circ$ from zenith for an off-centre source is $1.5\\sqrt{t}\\ \\sigma$. \\subsection{$\\gamma$-ray energy measurement} For a source detected with a strong enough significance, the energy spectrum can be studied by a detector with good energy resolution. The fit described above in which the source position is a free parameter gives a first estimate of the energy of each event to within about 30\\%. However, more precise spectral studies can be carried out on point sources of $\\gamma$-rays. The use of the source position as a constraint in the fit provides a higher accuracy for impact parameter measurement and, as a consequence, for energy measurement. If trigger selection effects are ignored in the Monte-Carlo program (thus accepting all events above $100\\:{\\mathrm {GeV}}$), the $\\chi^2$ minimization with respect to $E_\\gamma$, $D$, and $\\phi$ provides an unbiased energy measurement within about 25\\% (statistical error only). Trigger selection effects are small for events well above the threshold, as can be seen for simulated $400\\:{\\mathrm {GeV}}$ $\\gamma$-rays in Fig.~\\ref{fig:ener500}. \\begin{figure} \\epsfxsize=14.5 cm \\leavevmode \\centering \\epsffile[15 25 550 300]{fig12.eps} \\caption{Distribution of $\\ln({E_{\\mathrm {f}}}/{E_\\gamma})$ for vertical $400\\:{\\mathrm {GeV}}$ $\\gamma$-ray showers satisfying the selection cuts (the source location, $\\vec{\\xi}$, being fixed in the fit). The shaded histogram is further restricted to events with a fitted impact parameter $30\\:{\\mathrm {m}} < D_{\\mathrm {f}} < 125\\:{\\mathrm {m}}$.} \\label{fig:ener500} \\end{figure} This figure also shows that the distribution of the fitted event energies about the true energy is Gaussian on a logarithmic scale. Consequently, the slope of a power-law spectrum can be directly estimated with this technique. Close to the threshold, however, the fitted energy $E_{\\mathrm {f}}$ is overestimated as a consequence of the trigger selection. Similarly, the small remaining bias in $ \\log (E_{\\mathrm {f}}/E_\\gamma)$ at $400\\:{\\mathrm {GeV}}$ is due to showers with large impact parameters for which the trigger selection is critical at this energy since the telescope is then located at the border of the light pool (Fig.~\\ref{fig:densite}). This effect is largely removed if only showers with a fitted impact parameter $D_{\\mathrm {f}}$ lower than $125\\:{\\mathrm {m}}$ are included (shaded histogram in Fig.~\\ref{fig:ener500}). The bias induced by the trigger selection at different energies is best illustrated by plotting 68\\% confidence intervals for $E_{\\mathrm {f}}$ as a function of the true value $E_\\gamma$ used in the simulation (Fig.~\\ref{fig:enerfit}). \\begin{figure} \\epsfxsize=14.5 cm \\leavevmode \\centering \\epsffile[5 25 580 520]{fig14.eps} \\caption{Fitted energy, $E_{\\mathrm {f}}$, versus true energy, $E_\\gamma$, for $\\gamma$-ray showers satisfying the selection cuts (the source location, $\\vec{\\xi}$, being fixed in the fit) and for which $30\\:{\\mathrm {m}}< D_{\\mathrm {f}}\\cos Z < 125\\:{\\mathrm {m}}$ ($Z=$~zenith angle). The shaded interval shows the 68\\% confidence intervals for a source at the zenith. The effect of trigger selection on the energy estimation can be seen.} \\label{fig:enerfit} \\end{figure} It can be seen that for $E_f$ below $350\\:{\\mathrm {GeV}}$ for vertical showers, only an upper limit can be safely derived for $E_\\gamma$. Therefore, spectrum measurement is reliable only above a spectrometric threshold, that is, in the region in which $E_{\\mathrm {f}}$ depends linearly on $E$. It is somewhat higher than the nominal threshold which is relevant for source discovery. Fig.~\\ref{fig:enerfit} also shows the variation of the average values of $\\log(E_{\\mathrm {f}})$ on $\\log(E_\\gamma)$ for zenith angles $0^{\\circ}$, $30^{\\circ}$, $45^{\\circ}$, and $60^{\\circ}$, showing the increase in the spectrometric thresholds with increasing zenith angle. Restricting to events with $E_{\\mathrm {f}}$ above the spectrometric threshold and $30\\:{\\mathrm {m}}< D_{\\mathrm {f}} \\cos Z < 125\\:{\\mathrm {m}}$, the accuracy of the preceding method is about 20\\% (statistical error only), independent of the zenith angle $Z$ up to $45^{\\circ}$. ", "conclusions": "The method described above, based on a realistic analytic description of electromagnetic air showers, is best suited for those Cherenkov Imaging Telescopes with a high-resolution camera. The light distribution in the focal plane is fully exploited, yielding the shower direction from the asymmetry of the longitudinal profile as well as the source position in the focal plane. Selection of $\\gamma$-rays on the basis of the image shape is performed by using a single $\\chi^2$-variable instead of a series of cuts on various image parameters. By combining the $\\chi^2$ probability cut and a direction ($\\alpha$) cut, a significance of 5$\\sigma$ per hour can be achieved for a Crab-like source at the zenith. A future development of the method would be to use the distribution of selected $\\gamma$-ray origins on the celestial sphere together with the known energy-dependent point spread function of the method to estimate the significance by a maximum-likelihood method. With the C{\\small AT} telescope ($250\\:{\\mathrm {GeV}}$ threshold), sources with the intensity of the Crab nebula can be detected in one hour. This has been confirmed with the results obtained in the 96/97 observing campaign. Moreover, sources with poorly defined position can be localized with an accuracy of the order of an arc minute on the basis of about 100 showers. The accuracy on $\\gamma$-ray energy is of the order of 20--25\\%. Biases induced by the trigger selection have been investigated; in particular, care should be taken in spectrum measurement, which is accurate only above a specific threshold, higher than that used for source detection." }, "9804/astro-ph9804305_arXiv.txt": { "abstract": "We analyze the stability of g-modes in variable white dwarfs with hydrogen envelopes. All the relevant physical processes take place in the outer layer of hydrogen rich material which consists of a radiative layer overlain by a convective envelope. The radiative layer contributes to mode damping because its opacity decreases upon compression and the amplitude of the Lagrangian pressure perturbation increases outward. The convective envelope is the seat of mode excitation because it acts as an insulating blanket with respect to the perturbed flux that enters it from below. A crucial point is that the convective motions respond to the instantaneous pulsational state. Driving exceeds damping by as much as a factor of two provided $\\omega\\tau_c\\geq 1$, where $\\omega$ is the radian frequency of the mode and $\\tau_c\\approx 4\\tau_{\\rm th}$ with $\\tau_{\\rm th}$ being the thermal time constant evaluated at the base of the convective envelope. As a white dwarf cools, its convection zone deepens, and modes of lower frequency become overstable. However, the deeper convection zone impedes the passage of flux perturbations from the base of the convection zone to the photosphere. Thus the photometric variation of a mode with constant velocity amplitude decreases. These factors account for the observed trend that longer period modes are found in cooler DAVs. Overstable modes have growth rates of order $\\gamma\\sim 1/(n\\tau_\\omega)$, where $n$ is the mode's radial order and $\\tau_\\omega$ is the thermal time-scale evaluated at the top of the mode's cavity. The growth time, $\\gamma^{-1}$, ranges from hours for the longest period observed modes ($P\\approx 20$ minutes) to thousands of years for those of shortest period ($P\\approx 2 $ minutes). The linear growth time probably sets the time-scale for variations of mode amplitude and phase. This is consistent with observations showing that longer period modes are more variable than shorter period ones. Our investigation confirms many results obtained by Brickhill in his pioneering studies of ZZ Cetis. However, it suffers from at least two serious shortcomings. It is based on the quasiadiabatic approximation that strictly applies only in the limit $\\omega\\tau_c\\gg 1$, and it ignores damping associated with turbulent viscosity in the convection zone. We will remove these shortcomings in future papers. ", "introduction": "} ZZ Cetis, also called DAVs, are variable white dwarfs with hydrogen atmospheres. Their photometric variations are associated with nonradial gravity-modes (g-modes); for the first conclusive proof, see Robinson \\etal (\\cite{scaling-robinson82}). These stars have shallow surface convection zones overlying stably stratified interiors. As the result of gravitational settling, different elements are well separated . With increasing depth, the composition changes from hydrogen to helium, and then in most cases to a mixture of carbon and oxygen. From center to surface the luminosity is carried first by electron conduction, then by radiative diffusion, and finally by convection. Our aim is to describe the mechanism responsible for the overstability of g-modes in ZZ Ceti stars. This topic has received attention in the past. Initial calculations of overstable modes were presented in Dziembowski \\& Koester (\\cite{adia-dziem81}), Dolez \\& Vauclair (\\cite{adia-dolez81}), and Winget \\etal (\\cite{adia-winget82}). These were based on the assumption that the convective flux does not respond to pulsation; this is often referred to as the frozen convection hypothesis. Because hydrogen is partially ionized in the surface layers of ZZ Ceti stars, these workers attributed mode excitation to the $\\kappa$-mechanism. In so doing, they ignored the fact that the thermal time-scale in the layer of partial ionization is many orders of magnitude smaller than the periods of the overstable modes. Pesnell (\\cite{adia-pesnell87}) pointed out that in calculations such as those just referred to, mode excitation results from the outward decay of the perturbed radiative flux at the bottom of the convective envelope. He coined the term `convective blocking' for this excitation mechanism.\\footnote{This mechanism was described in a general way by Cox \\& Guili (\\cite{adia-cox68}), and explained in more detail by Goldreich \\& Keeley (\\cite{adia-goldreich77}).} Although convective blocking is responsible for mode excitation in the above cited references, it does not occur in the convective envelopes of ZZ Ceti stars. This is because the dynamic time-scale for convective readjustment (i.e., convective turn-over time) in these stars is much shorter than the g-mode periods. Noting this, Brickhill (\\cite{adia-brick83}, \\cite{adia-brick90}, \\cite{adia-brick91a}, \\cite{adia-brick91b}) assumed that convection responds instantaneously to the pulsational state. He demonstrated that this leads to a new type of mode excitation, which he referred to as convective driving. Brickhill went on to presents the first physically consistent calculations of mode overstability, mode visibility, and instability strip width. Our investigation supports most of his conclusions. Additional support for convective driving is provided by Gautschy \\etal (\\cite{adia-gautschy96}) who found overstable modes in calculations in which convection is modeled by hydrodynamic simulation. In this paper we elucidate the manner in which instantaneous convective adjustment promotes mode overstability. We adopt the quasiadiabatic approximation in the radiative interior. We also ignore the effects of turbulent viscosity in the convection zone. These simplifications enable us to keep our investigation analytical, although we appeal to numerically computed stellar models and eigenfunctions for guidance. The DA white dwarf models we use are those produced by Bradley (\\cite{scaling-bradley96}) for asteroseismology. Fully nonadiabatic results, which require numerical computations, will be reported in a subsequent paper. These modify the details, but not the principal conclusions arrived at in the present paper. The plan of our paper is as follows. The linearized wave equation is derived in \\S \\ref{sec:adia-prepare}. In \\S \\ref{sec:adia-perturb}, we evaluate the perturbations associated with a g-mode in different parts of the star. We devote \\S \\ref{sec:adia-driving} to the derivation of a simple overstability criterion. Relevant time-scales and the validity of the quasiadiabatic approximation are discussed in \\S \\ref{sec:adia-discussion}. The appendix contains derivations of convenient scaling relations for the dispersion relation, the WKB eigenfunction, and the amplitude normalization. ", "conclusions": "} \\subsection{Time-Scales \\label{sec:adia-relevant}} Three time-scales are relevant for convective driving in DAVs. The first is the period of an overstable g-mode, $P = 2\\pi/\\omega$, which is typically of order a few hundred seconds. The second is the dynamical time constant, $\\tcv \\sim H_p/\\vcv$, on which convective motions respond to perturbations; $\\tcv \\leq 1 \\s$ throughout the convection zones of even the coolest ZZ Cetis. This is why the convective motions adjust to the instantaneous pulsational state. The third is the thermal time constant, $\\tau_c$, during which the convection zone can bottle up flux perturbations that enter it from below. Given the central role of $\\tau_c$, we elaborate on its relation both to $\\tcv$ and to the more conventional definition of thermal time constant, $\\tau_{\\rm th}$, at depth $z$. The latter is the heat capacity of the material above that depth divided by the luminosity. In a plane parallel, fully ionized atmosphere this is equivalent to \\begin{equation} \\tau_{\\rm th}\\equiv {1\\over F} \\int_0^z dz \\, c_p\\, {\\rho k_B\\over m_p T} \\approx {5pz\\over 7F}. \\label{eq:adia-tauth}\\end{equation} Appeal to equation \\refnew{eq:adia-Fvcv} establishes that inside the convection zone \\begin{equation} {\\tcv\\over \\tau_{\\rm th}}\\sim \\left({\\vcv\\over c_s}\\right)^2\\ll 1. \\end{equation} Now $\\tau_c\\equiv (B+C)\\tau_b$, where $\\tau_b$ is defined by equation \\refnew{eq:adia-taub}. To the extent that $c_p\\approx 5$ is constant in the convection zone, $\\tau_b\\approx \\tau_{\\rm th}/5$, where the latter is evaluated at $z_b$. Next we address the relation between $\\tau_c$ and $\\tau_b$. Here we are concerned with the relatively large value of $B+C$, typically about 20 for DAVs.\\footnote{In our models, $B$ and $C$ have comparable value.} Recall from equations \\refnew{eq:adia-Dsb} and \\refnew{eq:adia-DFph} that \\begin{equation} {\\delta F_{\\ph} \\over F}\\approx {\\delta s_b\\over B+C}. \\label{eq:adia-dFds}\\end{equation} So the photosphere and superadiabatic layer add an insulating blanket on top of the convection zone. The large value of $B$ follows because the photospheres of DAVs are composed of lightly ionized hydrogen. In this state, the values of $\\kappa_T$ and $s_T$ are both large and positive; typical values in the middle of the instability strip are $\\kappa_T\\approx 6$ and $s_T\\approx 24$. The large and positive $\\kappa_T$ arises because the population of hydrogen atoms in excited states which the ambient radiation field can photoionize increases exponentially with increasing $T$. The large and positive $s_T$ occurs as the ionization fraction increases exponentially with increasing $T$, and the much larger entropy contributed by a free as compared to a bound electron. The large and positive value of $C$ reflects the increase in entropy gradient that accompanies an increase in convective flux. It is obtained from mixing length theory with an unperturbed mixing length. \\subsection{Validity of the Quasiadiabatic Approximation} The validity of the quasiadiabatic approximation requires that the nonadiabatic parts of the expressions for $\\delta \\rho/\\rho$ and $\\delta T/T$, as given by equations \\refnew{eq:adia-eqdelrho} and \\refnew{eq:adia-eqdelT}, be small in comparison to the adiabatic parts. Thus the ratio \\begin{equation} {\\cal R_{\\rm na}}\\equiv {\\delta s\\over\\delta p/p} \\label{eq:adia-rnonad}\\end{equation} is a quantitative measure of nonadiabaticity. We estimate ${\\cal R}_{\\rm na}$ for the radiative interior and the convection zone. We calculate $\\delta s$ in the radiative interior from \\begin{equation} \\delta s \\approx {iF\\over \\omega}{m_p\\over\\rho k_B T}{d\\over dz}\\left({\\delta F\\over F}\\right). \\label{eq:adia-delsrad}\\end{equation} In the upper evanescent layer, $z_bz_\\omega$, we find \\begin{equation} |{\\cal R_{\\rm na}}|\\sim {1\\over \\omega\\tau_{\\th}}\\left({z\\over z_\\omega}\\right). \\label{eq:adia-ratioevl}\\end{equation} Nonadiabatic effects in the radiative interior are maximal at $z=z_b$, where \\begin{equation} |{\\cal R_{\\rm na}}|\\sim {1\\over \\omega\\tau_b}, \\label{eq:adia-ratioevmax}\\end{equation} since $\\tau_{\\rm th}/z$ increases with depth. The measure of nonadiabaticity in the convection zone is given by equation \\refnew{eq:adia-Dsb}. Since $\\omega\\tau_b>1$ is required for the validity of the quasiadiabatic approximation in the radiative zone, we restrict consideration to the limiting case $\\omega\\tau_c\\gg 1$. In this limit, equation \\refnew{eq:adia-Dsb} yields \\begin{equation} |{\\cal R_{\\rm na}}|\\sim {1\\over \\omega\\tau_b}, \\label{eq:adia-ratiocvz}\\end{equation} which is identical to the value arrived at for the radiative zone in equation \\refnew{eq:adia-ratioevmax}. The requirement $\\omega\\tau_b\\gtrsim 1$ for the validity of the quasiadiabatic approximation severely limits the applicability of the current investigation. The perturbed flux at the photosphere is related to that at the bottom of the convection zone by \\begin{equation} {\\Delta F_{\\ph}\\over F}\\approx {1\\over 1-i\\omega\\tau_c}{\\Delta F_b\\over F}. \\label{eq:adia-relDelF}\\end{equation} Since $\\tau_c$ is at least an order of magnitude larger than $\\tau_b$, modes with $\\omega\\tau_b\\gtrsim 1$ are likely to exhibit small photometric variations. However, this may not render them undetectable because their horizontal velocity perturbations pass undiminished through the convection zone. \\subsection{Brickhill's Papers\\label{subsec:adia-brickhill}} Our investigation is closely related to studies of ZZ Cetis by Brickhill (\\cite{adia-brick83},\\cite{adia-brick90},\\cite{adia-brick91a}). Brickhill recognized that the convective flux must respond to the instantaneous pulsational state. To determine the manner in which the convection zone changes during a pulsational cycle, he compared equilibrium stellar models covering a narrow range of effective temperature. Brickhill provided a physical description of convective driving and obtained an overstability criterion equivalent to ours.\\footnote{Our time constant $\\tau_c$ is equivalent to the quantity $D$ which Brickhill defined in equation (9) of his 1983 paper.} Moreover, he recognized that the convection zone reduces the perturbed flux and delays its phase. Our excuses for revisiting this topic are that Brickhill's papers are not widely appreciated, that our approach is different from his, and that our paper provides the foundation for future papers which will examine issues beyond those he treated. \\begin{appendix}" }, "9804/astro-ph9804075_arXiv.txt": { "abstract": "The decay of massive neutrinos to final states containing only invisible particles is poorly constrained experimentally. In this letter we describe the constraints that can be put on neutrino mass and lifetime using CMBR measurements. We find that very tight lifetime limits on neutrinos in the mass range 10 eV - 100 keV can be derived using CMBR data from upcoming satellite measurements. \\\\ Keywords: Neutrino decay, Cosmology: Theory, Cosmic Microwave Background \\\\ PACS: 13.35.Hb, 14.60.St, 98.70.Vc ", "introduction": " ", "conclusions": "" }, "9804/astro-ph9804243_arXiv.txt": { "abstract": "Although the unified formula for $\\gamma$-ray absorption process involving both the magnetic field and a perpendicular electric field derived by Daugherty \\& Lerche (1975) is correct, we argued in this paper that their conclusion that the induced electric fields are important in the pair formation process in the pulsar magnetospheres is wrong and misleading. The key point is that usually the direction of a $\\gamma$ photon at the emission point observed in the laboratory frame should be $(v/c, 0, [1-(v/c)^2]^{1/2})$ rather than $(0, 0, 1)$, where $v$ is the co-rotating velocity. This emission direction is just the one which results in zero attenuation coefficient of the $\\gamma$ photon. Calculation shows that after the photon has moved a distance, its direction lead to the result that the induced electric field is also of minor importance. Thus only $\\gamma-B$ process is the important mechanism for the pair production in the pulsar magnetospheres. The implications of the modification by ejecting the induced electric field are also discussed. ", "introduction": "Pair production process plays an important role in pulsar physics. It is not only a necessary process for the multiplication of the particles to account for emissions of different bands from pulsars (e.g. Sturrock 1971; Ruderman \\& Sutherland 1975; Arons \\& Scharleman 1979; Arons 1983; Cheng, Ho, \\& Ruderman 1986), but also an important mechanism to absorb $\\gamma$-rays produced in the pulsar magnetospheres, especially near the polar cap region (e.g. Hardee 1977; Harding, Tademaru, \\& Esposito 1978; Harding 1981; Daugherty \\& Harding 1982, 1996; Zhao {\\it et al.} 1989; Lu \\& Shi 1990; Lu, Wei, \\& Song 1994; Dermer \\& Sturner 1994; Sturner, Dermer, \\& Michel 1995; Wei, Song, \\& Lu 1997). Furthermore, the way by which the $\\gamma$-rays are absorbed is also the key factor to limit the parameters of the inner magnetospheric accelerators of pulsars (e.g. Ruderman \\& Sutherland 1975, hereafter RS75; Zhang \\& Qiao 1996; Qiao \\& Zhang 1996; Zhang {\\it et al.} 1997a; Zhang, Qiao, \\& Han 1997b, hereafter ZQH97b). Pair formation in intense magnetic fields ($\\gamma-B$ process) has been studied explicitly by different authors (e.g. Erber 1966; Tsai \\& Erber 1974; Daugherty \\& Harding 1983, Rifert, M\\'{e}sz\\'{a}ros \\& Bagoly 1989), and its importance in pulsar physics was first pointed out by Sturrock (1971). Daugherty \\& Lerche (1975, hereafter DL75) first dealt with the case involving a relatively weaker electric field perpendicular to the magnetic field (${\\bf E}^2-{\\bf B}^2\\le0$, ${\\bf E\\cdot B}=0$), and came to a unified formula of the attenuation coefficient of the $\\gamma$ photons. The more general case involving both the perpendicular and the parallel components of the electric field with respect to the magnetic field ($E_\\perp$ and $E_\\parallel$) was presented by Daugherty \\& Lerche (1976) and Urrutia (1978). In the specific case of pulsars, although $E_\\parallel$ is usually sufficiently small so that its effect is negligible, $E_\\perp$ induced by the fast spin of the neutron stars was demonstrated to be very important in pair formation process by DL75. This leads many authors to take this effect seriously into account in their studies (e.g. Hardee 1977; Lu \\& Shi 1990; Lu, Wei, \\& Song 1994; Qiao \\& Zhang 1996). In this paper, we'll argue that although the unified formula of DL75 is correct, their conclusion that the induced electric fields are important in the pair formation process in the pulsar magnetospheres is wrong and misleading. The detailed argument is presented in Section 2 and Section 3. Finally, we discuss the possible implications of this modification. ", "conclusions": "Although the unified formula for the pair production process involving both the magnetic field and the perpendicular electric field derived by DL75 is correct, we have argued in this paper that their conclusion that the induced electric fields are important in the pair formation process in the pulsar magnetospheres is wrong and misleading at least for the ``aligned rotator'' case. At the emission point, the photon emitted by a certain mechanism (e.g. the curvature radiation or the inverse Compton scattering) just moves along the very direction in which the attenuation coefficient is zero. Considering the propagation of the photon, we found that this rotation-induced electric field still plays a minor role in the $\\gamma$-ray absorption process in the polar cap region of a pulsar. For the general case of an ``oblique rotator'' in which the magnetic and the rotational axes are misaligned, the co-rotating velocity is no more perpendicular to the magnetic field so that $\\theta_B^{\\prime}$ (also $\\theta_u^{\\prime}$ and $\\theta_\\gamma^{\\prime}$) is not $\\pi/2$. The photon direction consequently deviate from $(v/c, 0, [1-(v/c)^2]^{1/2})$ slightly. From Fig.2, we see that the attenuation coefficient is also very small around the direction $(v/c, 0, [1-(v/c)^2]^{1/2})$, so that the conclusion that the induced electric field plays a minor role in $\\gamma$-ray absorption still holds for the oblique rotator case. Actually, DL75's result can only be applied to the aligned case strictly, since generally the co-rotating velocity ${\\bf v}_{r}$ is not equal to ${\\bf v}_{drift}=c({\\bf E\\times B})/B^2$, with which one can define a frame where the electric field vanishes completely. In the co-rotating frame of an oblique rotator, an electric field component parallel to the magnetic field will still remain so that DL75's application condition fails. Our results in this paper may have some implications for some previous studies which regard the electric field as the important effect of $\\gamma$-ray absorption (e.g. Hardee 1977; Zhao {\\it et al.} 1989; Lu \\& Shi 1990; Lu, Wei, \\& Song 1994; Qiao \\& Zhang 1996). Although the polar cap models of the $\\gamma$-ray pulsars are by all means sound in principle, their concrete details will alter much by ejecting the induced electric field. Specifically, based on DL75's result, Hardee (1977) got an absolute upper limit to the photon energies $$E_\\gamma <9.6\\times 10^9 B_{12}^{-1}r_6^2 P {\\rm eV}, \\eqno(6)$$ (his Eq.(38)) at which escape of the $\\gamma$-rays from the magnetosphere is possible. It should be replaced by a threshold in the pure magnetic field absorption scheme $$E_\\gamma <2.6\\times 10^9 B_{12}^{-1} P^{1/2} {\\rm eV} \\eqno(7)$$ (Wei, Song, \\& Lu 1997, their Eq.(10), $r_6=1$ and the last open field line is assumed). Thus the generation order parameters proposed by Lu, Wei, \\& Song (1994) should take the form in Wei, Song, \\& Lu (1997, their Eq.(17)). The three boundary lines (birth line, death line and appearance line) in the $\\dot P-P$ diagram of pulsars derived by Qiao \\& Zhang (1996) are also based on the electric field absorption. The details will also be changed by ejecting the electric field, but the picture still remains and may give a hint to us about the magnetic field configuration in the neutron star vicinity (Qiao \\& Zhang, discussions)." }, "9804/astro-ph9804257_arXiv.txt": { "abstract": "Molecules dominate the cooling function of neutral metal-poor gas at high density. Observation of molecules at high redshift is thus an important tool toward understanding the physical conditions prevailing in collapsing gas. Up to now, detections are sparse because of small filling factor and/or sensitivity limitations. However, we are at an exciting time where new capabilities offer the propect of a systematic search either in absorption using the UV Lyman-Werner H$_2$ bands or in emission using the CO emission lines redshifted in the sub-millimeter. ", "introduction": "\\subsection{Introduction} QSO absorption line systems probe the baryonic matter over most of the history of the Universe (0~$<$~$z$~$<$~5). The so-called damped Ly$\\alpha$ (hereafter DLA) systems are characterized by a very large H~{\\sc i} column density ($N$(H~{\\sc i})~$>$~2$\\times$10$^{20}$ ~cm$^{-2}$), similar to what is usually observed through local spiral disks. The case for these systems to be produced by proto-galactic disks is supported by the fact that the cosmological density of gas associated with these systems is of the same order of magnitude as the cosmological density of stars at present epochs (Wolfe 1996). The presence of heavy elements ($Z \\sim 1/10 ~ Z_\\odot$) and the redshift evolution of metallicity suggest ongoing star formation activities in these systems (Lu et al. 1996, Pettini et al. 1996, 1997). Moreover, strong metal line systems have been demonstrated to be associated with galaxies at low and intermediate $z$ (e.g. Bergeron \\& Boiss\\'e 1991). It has also been shown that the profiles of the lines arising in the neutral gas show evidence for rotation (Wolfe 1996, Prochaska \\& Wolfe 1997). Whether these arguments are enough to demonstrate that DLA systems arise in large disks is a matter of debate however. Indeed simulations have shown that the progenitors of present day disks of galaxies could look like an aggregate of well separated dense clumps at high redshift. The kinematics could be explained by relative motions of the clumps with very little rotation (Haehnelt et al. 1997, Ledoux et al. 1998). Moreover, using {\\sl HST} high spatial resolution images of the field of seven quasars whose spectra contain DLA lines at intermediate redshifts (0.4~$<$~$z$~$<$~1), Le~Brun et al. (1997) show that, in all cases, at least one galaxy candidate is present within 4~arcsec from the quasar. There is no dominant morphological type in their sample: three candidates are spiral galaxies, three are compact objects and two are amorphous low surface brightness galaxies. Therefore, although the nature of the DLA systems is unclear they trace the densest regions of the Universe where star formation occurs.\\par \\subsection{Molecular hydrogen} \\begin{figure} \\epsfysize=15cm % \\epsfxsize=11cm % \\hspace{1.5cm}\\epsfbox{q0528mol.ps} % \\caption[h]{Fit result for a few rotational transitions of the H$_2$ Lyman absorption bands in the $z_{\\rm abs}$~=~2.8112 system toward PKS~0528--250. The spectrum has been obtained with the echelle spectrograph CASPEC attached on the ESO 3.6~m at La Silla. The resolution is $R$~=~36000 and the integration time 5~hours.} \\end{figure} It is thus surprising that despite intensive searches, the amount of H$_2$ molecules seems quite low in damped Ly$\\alpha$ systems in contrast to what is observed in our own galaxy. Two detections of H$_2$ molecules in high redshift DLA systems have been reported. Recently Ge \\& Bechtold (1997) have found strong absorptions in the $z_{\\rm abs}$~=~1.9731 DLA system toward Q~0013--004. They derive $N$(H$_2$)~=~6.9$\\times$ 10$^{19}$~cm$^{-2}$, $b$~=~15~km~s$^{-1}$, $T_{\\rm ex}$~$\\sim$~70~K and $n$(H)~$\\sim$~300~cm$^{-3}$ for a total hydrogen column density $N$(H)~=~6.4$\\times$10$^{20}$~cm$^{-2}$. This system has by far the largest H$_2$ abundance $f$~=~2$N$(H$_2$)/[2$N$(H$_2$)~+~$N$(H~{\\sc i})] $\\sim$~0.22$\\pm$0.05 observed in DLA systems. However the exact number should be confirmed using higher resolution data. Other searches have led to much smaller values or upper limits ($f$~$<$~10$^{-6}$, Black et al. 1987, Chaffee et al. 1988, Levshakov et al. 1992). Table~1 summarizes the caracteristics of damped Ly$\\alpha$ systems that have been searched for molecules. Levshakov \\& Varshalovich (1985) suggested that molecules could be present toward PKS~0528--250 at a redshift ($z_{\\rm abs}$~=~2.8112), slighly larger than the emission redshift of the quasar. This claim has been confirmed by Foltz et al. (1988) using a 1~\\AA~ resolution spectrum. The latter authors derive $N$(H$_2$)~=~10$^{18}$~cm$^{-2}$, $b$~=~5~km~s$^{-1}$, $T_{\\rm ex}$~=~100~K and log~$N$(H~{\\sc i})~=~21.1$\\pm$0.3. By fitting the damped absorption together with the Ly$\\alpha$ emission from the quasar, M\\o ller \\& Warren (1993) find log~$N$(H~{\\sc i})~=~21.35. New high resolution data has been recently obtained by Srianand \\& Petitjean (1998). They estimate the column density of H$_2$ molecules $N$(H$_2$)~$\\sim$~6$\\times$10$^{16}$~cm$^{-2}$ and the fractional abundance of H$_2$, $f$~=~5.4$\\times$10$^{-5}$ (see Fig.~1). The excitation temperature derived for different transitions suggests that the kinetic temperature of the cloud is $\\sim$200~K and the density $n$~$\\sim$~1000~cm$^{-3}$. The cloud has therefore a dimension along the line of sight smaller than 1~pc. Since it obscurs the broad-line emission region, its transverse dimension should be larger than 10~pc. Upper limits are obtained on the column densities of C~{\\sc i} ($<$~10$^{12.7}$~cm$^{-2}$) and CO ($<$~10$^{13.2}$~cm$^{-2}$; $N$(CO)/$N$(H~{\\sc i})~$<$~7$\\times$10$^{-9}$). It is suggested that the ratio $N$(H$_2$)/$N$(C~{\\sc i}) is a useful indicator of the physical conditions in the absorber. Photo-ionization models show that radiation fields with spectra similar to typical AGNs or starbursts are unable to reproduce all the constraints and in particular the surprizingly small $N$(C~{\\sc i})/$N$(H$_2$) and $N$(Mg~{\\sc i})/$N$(H$_2$) ratios. In view of the models explored, the most likely ionizing spectrum is a composite of a UV-\"big bump\" possibly produced by a local starburst and a power-law spectrum from the QSO that provides the X-rays. This suggests that the gas is not predominantly ionized by the quasar and that star-formation may occur in the clouds, a conclusion reached as well by Warren \\& M\\o ller (1996) and Ge et al. (1997). Dust is needed to explain the production of molecules in the cloud. The amount of dust is broadly consistent with the [Cr/Zn] abundance determination. \\par\\noindent \\vspace{0.5cm} % \\par\\noindent \\centerline{\\bf Tab. 1 - H$_2$ molecules in DLA systems} \\begin{table}[h] \\hspace{1.5cm} % \\begin{tabular}{|l|c|c|c|c|c|c|} \\hline Name & 000-263 & 0013-004 & 0100+130 & 0528-250 & 1331+170 & 1337+113 \\\\ \\hline $z_{\\rm em}$ & 4.110 & 2.084 & 2.681 & 2.770 & 2.081 & 2.919\\\\ $z_{\\rm abs}$ & 3.391 & 1.9731 & 2.309 & 2.811 & 1.776 & 2.796\\\\ $N$(HI) (10$^{21}$~cm$^{-2}$ & 2.0 & 0.64 & 2.5 & 2.2 & 1.5 & 0.80\\\\ $N$(H$_2$) (10$^{16}$~cm$^{-2}$& $<$~0.3 & 6900 & $<$~0.5 & 6 & ... & $<$~5\\\\ $f_{{\\rm H}2}$(10$^{-4}$) & $<$~0.03 & 2200 & $<$~0.04 & 0.5 & ... & $<$~1.3\\\\ \\hline \\multicolumn{7}{l}{Levshakov et al. (1992); Ge \\& Bechtold (1997); Srianand \\& Petitjean (1998)} \\\\ \\end{tabular} \\label{tsys} \\end{table} \\begin{figure} \\centerline{ \\psfig{figure=boisse.ps,width=9.cm,height=8.cm,angle=270} } \\caption[h]{ [Zn/H] versus log~$N$(H~{\\sc i}) for 37 damped Ly$\\alpha$ systems; small symbols correspond to $z_{\\rm abs}$~$<$~2.15 and large symbols to $z_{\\rm abs}$~$>$~2.15. The line in the upper right corresponds to $N$(Zn~{\\sc ii}) = 1.4$\\times$10$^{13}$~cm$^{-2}$ or to Galactic material inducing $A_{\\rm V}$~$\\sim$~0.27. The figure is taken from Boiss\\'e et al. (1998).} \\label{boisse} \\end{figure} \\subsection{Is there a bias against detection of H$_2$ molecules ?} The small number of H$_2$ detections in damped systems is intriguing. Indeed in the interstellar medium of our Galaxy, all the clouds with log~$N$(H~{\\sc i})~$>$~21 have log~$N$(H$_2$)~$>$~19 (Jenkins \\& Shaya 1979). Formation of H$_2$ is expected on the surface of dust grains if the gas is cool, dense and mostly neutral, and from the formation of negative hydrogen if the gas is warm and dust free (see e.g. Jenkins \\& Peimbert 1997). Destruction is mainly due to UV photons. The effective photodissociation of H$_2$ takes place in the energy range 11.1--13.6 eV, through Lyman-Werner band line absorption. In the DLA system toward PKS~0528-250, (i) abundances are of the order of 0.1~$Z_{\\odot}$; (ii) the ratio [Cr/Zn] indicates a depletion factor into dust-grains of the order of half of that in the Galactic ISM; (iii) although it has been shown that the cloud is located at a distance larger than 10~kpc from the quasar, it is still close to it and exposed to its UV flux. Nonetheless, molecular hydrogen is detected. This suggests that indeed, molecular hydrogen should be seen in most of the damped systems. The small number of detections may be explained if observations are biased against the presence of molecules. Indeed it can be speculated that molecules should be found predominantly in gas with a non negligible amount of dust. However the corresponding extinction of the background quasar due to the dust in the damped system could be large enough to drop the quasar out of the sample of quasars that are usually observed for such studies. Boiss\\'e et al. (1998) notice that for the damped systems studied up to now, the larger the H~{\\sc i} column density, the smaller the abundances (see Fig.~\\ref{boisse}). This suggests that the high column density DLA systems detected up to now are those with the smallest metallicities and consequently those with the smallest amount of dust. One way to clear up this problem is to observe a complete sample of quasars (if possible constructed without color selection) and to search for high column density damped systems. ", "conclusions": "" }, "9804/astro-ph9804061_arXiv.txt": { "abstract": "We have computed stellar evolutionary models for stars in a mass range characteristic of Cepheid variables ($3 10000$ gravothermal oscillations are clearly visible. The calculations of $N=2000$, $4096$, $10000$, $32000$ and $100000$ models take about $2$, $6$ $20$, $130$ and $2500$ hours, respectively. The Monte--Carlo code is at least $10^5$ times faster than the $N$--body one for $N=32768$ with special--purpose hardware (Makino 1996ab). Thus it becomes possible to run several different models to improve statistical quality of the data and run individual models with $N$ as large as $100000$. The Monte--Carlo scheme can be regarded as a method which lies in the middle between direct $N$--body and Fokker--Planck models and combines most advantages of both methods. ", "introduction": "Our knowledge about the stellar content, kinematics, and the influence of the environment on observational features of globular clusters and even richer stellar systems are increasing dramatically (Janes 1991, Djorgovski \\& Meylan 1993, Smith \\& Brodie 1993, Hut \\& Makino 1996, Meylan \\& Heggie 1997). First, observations are reaching the point where segregation of mass within globular clusters can be observed directly and quantitatively. Second, observations have revealed that clusters with dense (collapsed) cores are relatively more concentrated to the galactic center than uncollapsed ones. Thus the influences of the environment and mass spectrum are crucial for cluster evolution. Third, observations give clear evidence that post--collapse globular clusters have bluer cores. This suggests strong influence of dynamical interactions between stars on observational properties of globular clusters. Fourth, recent observations show that many different and fascinating types of binaries and binary remnants are present in abundance in globular clusters. Binaries, in addition to being a diagnostic of the evolutionary status of clusters, are directly involved in the physical processes of energy generation, providing the energy source necessary to stop the core collapse and then drive the core expansion. So, to model the evolution of real stellar systems and make meaningful comparison with observation one has to take into account the complex interactions between stellar evolution, stellar dynamics and the environment. Of course all these demands can be fulfilled by direct $N$--body codes (but even the $N$--body method will have trouble with stellar evolution of binary stars). But they are very time--consuming and they need a special--purpose hardware to be run efficiently (Makino 1996ab). Another possibility is to use a code which is very fast and properly reproduces the standard relaxation process and at the same time provides a clear and unambiguous way of introducing all the physical processes which are important during globular cluster evolution. This task might seem unachievable, but actually this kind of code was in use in the past. Monte--Carlo codes, which use a statistical method of solving the Fokker--Planck equation provide all the necessary flexibility. They were developed by Spitzer (1975, and references therein) and H\\'enon (1975, and references therein) in the early seventies, and substantially improved by Marchant \\& Shapiro (1980, and references therein) and Stod\\'o\\l kiewicz (1986a, and references therein). Unfortunately, lack of fast computers with sufficient memory at that time and development of the direct Fokker--Planck and gaseous models contribute to the abandonment of this method. But recent developments in computer hardware, speed and memory now make it possible to run a Monte--Carlo code efficiently, even on general--purpose workstations. The great advantages of this method, beside of its simplicity and speed, are connected with the inclusion of anisotropy and with the fact that added realism does not slow it down. The Monte--Carlo method can practically cope as easily as the $N$--body method with internal freedom of single and binary stars and external environment, with one exception, a stellar system must be spherically symmetric. The Monte--Carlo code can have another possible use. Despite the simplified nature of continuum models (Fokker--Planck and gaseous models) they will continue for a while to be the most commonly used codes for stellar dynamical evolution. The Monte--Carlo models can be used to optimise physical free parameters and approximations of continuum models to check their validity as it was done in comparison between small $N$--body simulations and continuum ones (Giersz \\& Heggie 1994ab, Giersz \\& Spurzem 1994). This procedure should further increase our confidence in results obtained by Fokker--Planck or gaseous simulations. On the other hand the Monte--Carlo techniques can be incorporated in continuum models to describe the stochastic processes of binary formation, energy generation and movement (Spurzem \\& Giersz 1996, Giersz \\& Spurzem 1997). This, for example, will enable a very detailed investigation of evolution of primordial binaries in evolving background given by an anisotropic gaseous model. The plan of the paper is as follows. In Section 2 a short review of the `old' and `new' Monte--Carlo methods will be presented. In Section 3 the first results of the `new' Monte--Carlo simulation will be presented. And finally in Section 4 the conclusions and future development of the code will be discussed. ", "conclusions": "\\medskip A successful revision of Stod{\\'o}{\\l}kiewicz's Monte--Carlo code was presented. The updated method treats each {\\it superstar} as a single star and follows the evolution and motion of all individual stellar objects. This improvement was possible thanks to the recent developments in computer hardware and computer speed. Two essential changes was added to the original Monte--Carlo code. Firstly, the procedure which deal with problems of radial velocity determination after the system rearrangement (changes of mechanical energy of the stars due to changes of mass distribution) was slightly changed. This assures better energy conservation. Secondly, the new procedure which deals with star escapers was added. This practically resolves the problem with too high escape rate observed in Monte--Carlo simulations. The Monte--Carlo scheme presented here (as previous Monte--Carlo schemes) takes full advantage of the undisputed physical knowledge on the secular evolution of (spherical) star clusters as inferred from continuum model simulations. Additionally it describes in a proper way the graininess of the gravitational field and the stochasticity of the real $N$--body systems. This does not include any additional physical approximations or assumptions which are common in Fokker--Planck and gas models (e.g. conductivity or isotropic distribution function for field stars). From that respect Monte--Carlo scheme can be regarded as a method which lies in the middle between direct $N$--body and Fokker--Planck models and combines most advantages of the both methods. The first calculations for equal--mass $N$--body systems with three--body energy generation according to Spitzer's formulae show good agreement with direct $N$--body calculations for $N=2000$, $4096$ and $10000$ particles. The density, velocity, mass distributions, energy generation, number of binaries etc. follow the $N$--body results. Only the number of escapers is slightly too high compared to $N$--body results (but this can be resolved by the time--dependent shift of the escape rate) and there is no level off anisotropy for advanced post--collapse evolution of Monte--Carlo models as is seen in $N$--body simulations for $N \\leq 2000$. For simulations with $N > 10000$ gravothermal oscillations are clearly visible. This is the first unambiguous detection of gravothermal oscillations in Monte--Carlo simulations. Moreover, this is a first unambiguous detection of gravothermal oscillations for stochastic $N$--body system with $N$ as large as $100000$. The speed of the new code makes it possible to run individual models with $N$ as large as $100000$ and also enables, in an unambiguous way, the inclusion of several different physical processes which operate during different stages of evolution of real globular clusters. The new Monte--Carlo code described in this paper is seen as a first step towards realistic models of globular clusters. Several important physical processes have to be included to make the simulations of the stellar systems more realistic. The final code will contain the following physical processes: {\\bf (1)} formation of binaries due to dynamical and tidal interactions, {\\bf (2)} primordial binaries, {\\bf (3)} stellar evolution, {\\bf (4)} tidal field of Galaxy and tidal shocks connected with crossing the galactic plane and with large molecular clouds, {\\bf (5)} collisions between stars, {\\bf (6)} interactions between binaries and stars and between binaries themselves, improving the presently used scattering cross-sections for binary hardening. In the first stage all processes connected with interactions between objects were modelled using analytical cross sections available in the literature. This allowed the code to be tested, and made possible comparison with continuum models. In the next stage interactions between groups of three and four stars will be modelled by numerical integrations of their orbits (the first attempts are tested now). If during the integration the distance between two or more stars becomes smaller than the sum of their radii then a physical collision takes place. This more realistic approach ensures that processes of energy generation (the most important factor in the dynamical evolution of globular clusters) will be modelled more closely. The final stage will be the inclusion of detailed 3--D hydrodynamical modelling of collisions between stars. This will be done by use of Smooth Particle Hydrodynamics (SPH) for a limited number of particles per star (a few hundred). This will allow close comparison between numerical models and observations of real globular clusters. I refer here to observations of various, peculiar objects like blue stragglers and milliseconds pulsars, which can be formed during collisions and encounters between stars. \\bigskip \\bigskip {\\parindent=0pt {\\bf Acknowledgments} I would like to thank Douglas C. Heggie and Rainer Spurzem for stimulating discussions, comments and suggestions to a draft version of this paper. I also thank Douglas C. Heggie, who made the $N$--body results for $N = 4096$ particles available. This work was supported in part by the Polish National Committee for Scientific Research under grant 2--P304--009-06.}" }, "9804/astro-ph9804194_arXiv.txt": { "abstract": "We present a Hubble Space Telescope (HST) study of the nuclear region of the E4 radio galaxy NGC 7052, which has a nuclear disk of dust and gas. The Second Wide Field and Planetary Camera (WFPC2) was used to obtain {\\BB}, {\\VV} and~{\\II} broad-band images and an {\\halnii} narrow-band image. The images yield the stellar surface brightness profile, the optical depth of the dust, and the flux distribution of the ionized gas. The Faint Object Spectrograph (FOS) was used to obtain {\\halnii} spectra at six different positions along the major axis, using a $0.26''$ diameter circular aperture. The emission lines yield the rotation curve of the ionized gas and the radial profile of its velocity dispersion. The observed rotation velocity at $r=0.2''$ from the nucleus is $V = 155 \\pm 17 \\kms$. The Gaussian dispersion of the emission lines increases from $\\sigma \\approx 70\\kms$ at $r=1''$, to $\\sigma \\approx 400\\kms$ on the nucleus. To interpret the gas kinematics we construct axisymmetric models in which the gas and dust reside in a disk in the equatorial plane of the stellar body, and are viewed at an inclination of $70^{\\circ}$. It is assumed that the gas moves on circular orbits, with an intrinsic velocity dispersion due to turbulence (or otherwise non-gravitational motion). The latter is required to fit the observed increase in the line widths towards the nucleus, and must reach a value in excess of $500 \\kms$ in the central $0.1''$. The circular velocity is calculated from the combined gravitational potential of the stars and a possible nuclear black hole. Models without a black hole predict a rotation curve that is shallower than observed ($V_{\\rm pred} = 92\\kms$ at $r=0.2''$), and are ruled out at $>99$\\% confidence. Models with a black hole of mass $\\Mbh = 3.3^{+2.3}_{-1.3} \\times 10^8 \\Msun$ provide an acceptable fit. The best-fitting model with a black hole adequately reproduces the observed emission line shapes on the nucleus, which have a narrower peak and broader wings than a Gaussian. NGC 7052 can be added to the list of active galaxies for which HST spectra of a nuclear gas disk provide evidence for the presence of a central black hole. The black hole masses inferred for M87, M84, NGC 6251, NGC 4261 and NGC 7052 span a range of a factor $10$, with NGC 7052 falling on the low end. By contrast, the luminosities of these galaxies are identical to within $\\sim\\!25$\\%. Any relation between black hole mass and luminosity, as suggested by independent arguments, must therefore have a scatter of at least a factor $10$. ", "introduction": "Astronomers have been searching for direct evidence for the presence of black holes (BHs) in galactic nuclei for more than two decades. Initially, the only constraints on the central mass distributions of galaxies were obtained from ground-based stellar kinematical observations. More recently, the launch of the Hubble Space Telescope (HST) and the subsequent refurbishment in 1993 have provided an important increase in spatial resolution. Combined with new techniques for data analysis and dynamical modeling this has strengthened the stellar kinematical evidence for BHs in several quiescent galaxies (e.g., Kormendy \\etal 1996a,b; van der Marel \\etal 1997a; Cretton \\& van den Bosch 1998; Gebhardt \\etal 1998). New tools for the detection of BHs were also developed. HST observations of the rotation velocities of nuclear disks of ionized gas provided accurate BH mass determinations for several active galaxies (e.g., Harms \\etal 1994; Ferrarese, Ford \\& Jaffe 1996, 1998; Macchetto \\etal 1997; Bower \\etal 1998), while for other galaxies BHs were detected through VLBI observations of nuclear water maser sources (e.g., Miyoshi \\etal 1995). The case for a BH in our own galaxy improved drastically through measurements of stellar proper motions exceeding $1000 \\kms$ in the central $0.1\\pc$ (Genzel \\etal 1997). There are now a total of 10---20 galaxies for which a nuclear dark mass, most likely a BH, has been convincingly detected. The combined results for these galaxies are summarized and reviewed in, e.g., Kormendy \\& Richstone (1995), Ford \\etal (1998), Ho (1998), Richstone (1998), and van der Marel (1998). This sample is now large enough to study the BH mass distribution in galaxies, which is further constrained by ground-based stellar kinematical observations (Magorrian \\etal 1998), HST photometry (van der Marel 1998) and quasar evolution (e.g., Haehnelt \\etal 1998). Our understanding remains sketchy, but is consistent with a picture in which a majority of galaxies has BHs, and in which the BH mass $\\Mbh$ correlates with the luminosity or mass of the host spheroid. In this paper we present and analyze HST data for the E4 galaxy NGC 7052. This galaxy is a radio source with a core and jet, but no lobes (Morganti \\etal 1987). Ground-based optical images show a nuclear dust disk aligned with the major axis of the galaxy (Nieto \\etal 1990). The physical properties of this disk were discussed by de Juan, Colina \\& Golombek (1996). In a previous paper (van den Bosch \\& van der Marel 1995; hereafter Paper~I) we presented ground-based narrow-band imaging and long-slit spectroscopy obtained with the 4.2m William Herschel Telescope (WHT). These observations showed that there is also a rotating nuclear disk of ionized gas in NGC 7052. The gas has a steep central rotation curve, rising to nearly $300 \\kms$ at $1''$ from the center. However, the spatial resolution of the spectra was insufficient to convincingly detect a BH, due in part to the relatively large distance of NGC 7052 ($58.7 \\Mpc$; i.e., $1'' = 284.6 \\pc$). The velocity dispersion of the gas was found to increase from $\\sigma \\approx 70\\kms$ at $1''$ from the center to $\\sigma \\approx 200\\kms$ on the nucleus. We showed that this cannot be the sole result of rotational broadening, which would have predicted double-peaked line profile shapes that are not observed. Instead, the observed central increase in the line width must be at least partly intrinsic. The ground-based kinematics yield an upper limit of $\\sim 10^9 \\Msun$ on the mass of any possible BH. To improve the constraints on the presence of a central BH we obtained broad- and narrow-band images of NGC 7052 with the Second Wide Field and Planetary Camera (WFPC2) and spectroscopy with the Faint Object Spectrograph (FOS), both in the context of HST project GO-5848. We discuss the imaging and photometric analysis in Section~\\ref{s:WF}, and the spectroscopy and kinematical analysis in Section~\\ref{s:FOS}. In Section~\\ref{s:dyn} we construct dynamical models to interpret the results. With the high spatial resolution of these data we are able to better constrain the nuclear mass distribution, and we find that NGC 7052 has a BH with mass $\\Mbh = 3.3^{+2.3}_{-1.3} \\times 10^8 \\Msun$. We summarize and discuss our findings in Section~\\ref{s:disc}. Some observational details are presented in an appendix. We adopt $H_0 = 80 \\kms \\Mpc^{-1}$ throughout this paper. This does not directly influence the data-model comparison for any of our models, but does set the length, mass and luminosity scales of the models in physical units. Specifically, distances, lengths and masses scale as $H_0^{-1}$, while mass-to-light ratios scale as $H_0$. ", "conclusions": "\\label{s:disc} We have presented HST observations of the nuclear gas and dust disk in the E4 radio galaxy NGC 7052. WFPC2 broad- and narrow-band images were used to constrain the stellar surface brightness profile, the optical depth of the dust, and the flux distribution of the ionized gas. We have built axisymmetric models in which the gas and dust reside in the equatorial plane, and in which the gas moves on circular orbits with an additional velocity dispersion due to turbulence (or otherwise non-gravitational motion). These models were used to interpret the ionized gas kinematics inferred from our new FOS spectra and from existing ground-based spectra. The models fit the observed central rotation gradient only if there is a central BH with mass $\\Mbh = 3.3^{+2.3}_{-1.3} \\times 10^8 \\Msun$. Models without a black hole are ruled out at $>99$\\% confidence. The models provide an adequate fit to the available observations with a minimum number of free parameters. The assumptions that we make are similar to those that have been made in HST studies of other galaxies with nuclear gas disks. In several areas our models are in fact more sophisticated than some of the previous work. In particular: we use our multi-colour photometry in order to constrain the central cusp steepness of the stellar mass distribution; we explicitly take into account the contribution of the axisymmetric stellar mass distribution to the circular velocity of the gas, and we do not assume the rotation field to be purely Keplerian; we explicitly model the convolution with the HST/FOS PSF and the binning over the size of the aperture; we model the full line profile shapes, and fit the widths of the emission lines as well as their mean; and we fit Gaussians to the models as we do the data, to properly take into account the fact that Gaussian fits to lines that may be skewed or have broad wings yield biased estimate of the true moments. Still, our models remain only an approximation to the true structure of NGC 7052. In particular: the thickness of the gas disk may not be negligible; the mean motion of the gas may not be circular; and the observed rotation curve may not perfectly reflect the intrinsic rotation curve, because of partial absorption of the emission line flux by dust. The limited sky coverage of the FOS spectra prevents a direct check on whether the gas motions in NGC 7052 are indeed circular. However, several consistency checks are available that may have signaled errors in our assumptions; none did. The stellar mass-to-light ratio and systemic velocity inferred with our models from the nuclear gas kinematics agree with those inferred from stellar kinematical measurements outside the region influenced by dust absorption. The best-fitting model for the gas kinematics reproduces the shapes of the emission lines on the nucleus, despite the fact that these shapes were not included as constraints in the fit. These agreements do not rule out a conspiracy of some sort, but they do make it less likely that the observed gas kinematics are the result of vastly non-circular motion, or have been strongly modified by dust absorption. Models of adiabatic BH growth for the stellar surface brightness cusp provide another successful check: the BH mass implied by these models is fully consistent with that inferred from the gas kinematics. Figure~\\ref{f:allBHs} shows a scatter plot of $\\Mbh$ versus {\\BB}-band spheroid luminosity $L_{B,{\\rm sph}}$ for all galaxies with reasonably secure BH mass determinations (adapted from van der Marel 1998, with the addition of NGC 7052; all for $H_0 = 80 \\kms \\Mpc^{-1}$). There is a trend of increasing $\\Mbh$ with increasing $L_{B,{\\rm sph}}$, although it remains difficult to rule out that systematic biases play some role in this relation (van der Marel 1998). Besides NGC 7052, the other galaxies for which the BH detections are based on kinematical studies of nuclear gas disks with the HST are M87 (Harms \\etal 1994; Macchetto \\etal 1997), M84 (Bower \\etal 1998), NGC 6251 and NGC 4261 (Ferrarese, Ford \\& Jaffe, 1998, 1996). The $\\Mbh$ in these galaxies are $3.2 \\times 10^9$, $1.4 \\times 10^9$, $6.6 \\times 10^8$ and $\\Mbh = 4.9 \\times 10^8 \\Msun$, respectively. NGC 7052 falls at the low end of this range. The five galaxies with BH evidence from nuclear gas disks form a very homogeneous set. Each of these galaxies is a radio source and is morphologically classified as an elliptical. The luminosities are identical to within $\\sim\\!25$\\% ($\\log L_B$ in the range $10.6$---$10.8$ for all five galaxies). By contrast, the black hole masses span a range of a factor $10$. The results for these galaxies therefore show that any relation between $\\Mbh$ and $L_{B,{\\rm sph}}$ must have a scatter of at least a factor $10$, even if the comparison is restricted to galaxies of similar type. \\placefigure{f:allBHs}" }, "9804/astro-ph9804037_arXiv.txt": { "abstract": "Two dimensional realizations of self-consistent models for the ``perfect elliptic disks'' were tested for global stability by gravitational N-body integration. The family of perfect elliptic disk potentials have two isolating integrals; time independent distribution functions $f(E,I_2)$ which self-consistently reproduce the density distribution can be found numerically, using a modified marching scheme to compute the relative contributions of each member in a library of orbits. The possible solutions are not unique: for a given ellipticity, the models can have a range of angular momenta. Here results are presented for cases with minimal angular momentum, hence maximal random motion. As in previous work, N-body realizations were constructed using a modified quiet start technique to place particles on these orbits uniformly in action-angle space, making the initial conditions as smooth as possible. The most elliptical models initially showed bending instabilities; by the end of the run they had become slightly rounder. The most nearly axisymmetric models tended to become more elongated, reminiscent of the radial orbit instability in spherical systems. Between these extremes, there is a range of axial ratios $0.305 \\lesssim b/a \\lesssim 0.570$ for which the minimum streaming models appear to be stable. ", "introduction": "Recent studies of elliptical galaxies and of bulges of spiral galaxies indicate that their figures are likely to be at least slightly triaxial (for reviews see \\cite{bin82}; \\cite{dzf91}; \\cite{bs93}). Most elliptical galaxies appear to be supported at least in part by anisotropies in the velocity distributions rather than by rapid rotation: see, for example, the work on the dwarf elliptical galaxies NGC 147, 185 and 205 by \\cite{bpn91} and \\cite{hdmp92}. A class of non-rotating potentials, known as the perfect ellipsoids, has been advanced as a possible model for elliptical galaxies (e.g. \\cite{dez85}). In these potentials, the mass density is stratified on concentric, similar ellipsoids, and is non-singular in the center. Many of the properties of these potentials can be derived analytically; the orbits all have three isolating integrals, and hence properties such as the time-averaged density distribution can be computed exactly. This simplifies the task of finding {\\it self-consistent models\\/}: time-steady phase-space distribution functions $f({\\bf x}, {\\bf v})$ such that the resulting mass density generates the desired gravitational potential. \\cite{stat87} and \\cite{teub87} have demonstrated that distribution functions for the perfect ellipsoids, and the analogous two-dimensional elliptic disks, can be constructed. Various sub-families of the axisymmetric perfect ellipsoids have been tested for stability (\\cite{dzs89}; \\cite{ms90}; \\cite{mh91}; \\cite{rdz91}). Flattened perfect ellipsoids could also be viewed as models for galactic bars. The only analytical bar models are Freeman's (\\cite{free66a}, \\cite{free66b}, \\cite{free66c}) bars, which are based upon a rotating two dimensional harmonic oscillator potential, and the perfect elliptic disk models, which have no figure rotation. \\cite{tdz87} showed that in the limit of the needle ($b \\rightarrow 0$) the two dimensional perfect elliptic disk is neutrally stable. Prompted by the large streaming velocities seen in barred spiral galaxies, the stability of perfect elliptic disks with maximum angular momentum has already been studied (\\cite{slls94}). The roundest disks were unstable to spiral mode formation, the most elongated elliptical models were unstable to bending modes, while the models with axial ratio $b/a$ in the range $0.250 \\lesssim b/a \\lesssim 0.570$ appeared stable. This paper extends that previous work to the study of a set of low angular momentum perfect elliptical disks. The minimal angular momentum cases allow us to study the ability of internal velocity dispersion to support an elliptic figure, and forms a natural complement to the earlier work as the other bound of the whole class of perfect elliptic disks. As before, we tested for global stability by constructing a discrete, self-consistent model, loading it into an $N$-body integrator and allowing it to evolve. ", "conclusions": "In this paper and in LS, we have constructed discrete self-consistent representations of the distribution functions of a range of perfect elliptic disks with minimal and maximal angular momentum. These models were then integrated forward in time using an $N$-body integrator to see if they were stable. The nearly axisymmetric and the most elongated models were unstable. The perfect elliptic disks with moderate axial ratios appear to be stable in both the maximum and minimum streaming cases. In the maximum streaming case, the nearly axisymmetric models developed spiral and bar instabilities as expected, since their limiting case, a cold axisymmetric disk, is known to be violently unstable to spiral instabilities. In the minimal angular momentum case, nearly-round disks became more elliptical, in a manner very similar to the radial orbit instability of spherical systems. This is not too surprising, since the velocity distribution is anisotropic, with the radial velocity dispersion being substantially higher than the tangential dispersion, even in the very nearly axisymmetric models. This comes about because of the substantial presence of box and marginal loop orbits in the models. In both angular momentum extremes, the most elongated models developed a bending instability. The similarity in behavior is not very surprising given the decreasing importance of rotational support with increasing ellipticity in these models. \\cite{tdz87} have shown that the limiting case of the needle ($b\\rightarrow0$) is neutrally stable to bending, while Merritt \\& Hernquist (1991)\\nocite{mh91} have demonstrated a bending instability in a very prolate (E9) system. It is thus not surprising that the most elliptic models should develop this instability. For the minimum angular momentum family, the instabilities change the shape of the disk towards a more moderate ellipticity. In the nearly axisymmetric disks, as the angular momentum is decreased from a maximum, we expect that the increasing velocity dispersion should help to stabilize against spiral instabilities. It appears likely that there is a stable region for nearly axisymmetric disks with values of Toomre's (\\cite{t64}) stability parameter $Q$ which lies between the points $Q \\sim 2$ and $Q \\sim 3$ where the velocity dispersion has increased to the point of being able to support the disk against the spiral instability (fig \\ref{fig-angmom}). As the rotational support becomes negligible and the radial velocity dispersion increases, a radial--orbit instability develops; the disks with lower angular momentum become unstable to elliptical distortions when $T_{\\rm radial} / T_{\\rm tangential} \\gtrsim 1.2$ (after the discussion of \\cite{fp84} and \\cite{p87} for stability of spherical systems). We expect that there is a range of angular momentum between the two extremes for which the nearly round disks are stable. The moderately elliptical disks with maximum and minimum angular momentum appear to be stable, so we would anticipate that disks of similar ellipticity and intermediate angular momentum will also be stable. The stability of the two-dimensional models with moderate ellipticity gives us hope that the three dimensional perfect ellipsoids of intermediate triaxiality (which is probably the appropriate range for elliptical galaxies \\cite{mb81}; \\cite{ddc76}; \\cite{dzf91}), will also be stable. It is known that some very flattened systems, such as the extreme oblate spheroids constructed from thin short-axis tube orbits (\\cite{ms90}), are unstable, but the simple fact that two longer axes are unequal is not likely to be the cause of further trouble. The three dimensional extension of this work will be interesting to see in light of the work of \\cite{app92} showing that three dimensional systems with a small amount of rotational streaming are unstable to a tumbling bar instability, both when the models have largely radial orbits and when the orbits are mostly circular. The techniques developed in this work have laid the foundation for investigating the stability of three dimensional perfect ellipsoids, and indeed of any integrable potential. The methods for choosing orbits, whether simulated annealing (as in LS) or the marching scheme of ZHS, can be easily expanded to take account of a variety of possible cost terms related to angular momentum or line of sight velocities (e.g. \\cite{rix97}). The procedure of LS for generating a quiet start which minimizes random noise due to particle discreteness, can be carried over to any integrable system. This is potentially most useful in $N$-body studies which attempt to measure the growth rate of instabilities, because the detection of instabilities which are still in the linear regime is limited by particle noise. For example, \\cite{s91} found that his linear stability theory was consistent with the results of $N$-body simulations for highly unstable spherical systems, but predicted slow growing instabilities which could not be seen in the simulations because of particle noise. \\cite{app90} have constructed an analytic potential--smoothing integration technique which decreases the $\\sqrt{N}$ noise associated with binning and softening in $N$-body codes, and permits better examination of the linear growth regime. Their method would also benefit from a quiet start, because the particle discreteness then makes a larger relative contribution to the noise." }, "9804/astro-ph9804171_arXiv.txt": { "abstract": "In this letter we present an idea which reconciles a homogeneous and isotropic Friedmann universe with a fractal distribution of galaxies. We use two observational facts: The flat rotation curves of galaxies and the (still debated) fractal distribution of galaxies with fractal dimension $D=2$. Our idea can also be interpreted as a redefinition of the notion of bias.\\\\ {\\bf Key Words:} Large scale structure of universe-cosmology: theory-cosmology:dark matter galaxies:general ", "introduction": " ", "conclusions": "" }, "9804/hep-ph9804444_arXiv.txt": { "abstract": "The process of photon splitting $\\gamma \\to \\gamma \\gamma$ in a strong magnetic field is investigated both below and above the pair creation threshold. Contrary to the statement by Baier et al., the ``allowed'' channel $\\alw$ is shown not to be a comprehensive description of splitting in the strong field because the ``forbidden'' channel $\\frb$ is also essential. The partial amplitudes and the splitting probabilities are calculated taking account of the photon dispersion and large radiative corrections near the resonance. \\\\\\\\ PACS numbers: 12.20.Ds, 95.30.Cq, 98.70.Rz ", "introduction": " ", "conclusions": "" }, "9804/astro-ph9804165_arXiv.txt": { "abstract": "Images and longslit, echelle spectra of the \\Ha emission from 14 dwarf galaxies and M82 have been used to identify expanding shells of ionized gas. Supershells (radius $>~300$~pc) are found in 12 of the dwarfs. The measured shell sizes and expansion speeds constrain the ages and power requirements of the bubbles. The dynamical age of the larger bubbles is typically about 10~Myr, and ionized shells older than 20~Myr are rare. An energy equivalent to 100 to 10,000 supernova explosions over this period is needed to drive the shock front that sweeps out the cavity. The current star formation rates are high enough to meet these power requirements. Many of the shells will breakthrough the surrounding layer of HI supersonically, but the projected expansion speeds are typically less than the lower limits on the escape velocity. Some of the shell material may permanently escape from a few galaxies such as \\n1569. Whether bound to the galaxy or not, these outflows probably play an important role in regulating the star formation rate and are expected to significantly influence the chemical evolution of the galaxies. The shells lift gas out of the disk at rates comparable to, or even greater than, the current galactic star formation rates. They will only displace a substantial fraction of the interstellar gas if their duty cycle is much longer than the rotational period of the disk. ", "introduction": "The interplay between massive stars and the interstellar medium (ISM) plays a fundamental role in the formation and evolution of galaxies. In addition to ionizing radiation and newly synthesized elements, massive stars deliver kinetic energy and momentum to the surrounding gas through stellar winds and supernova explosions. Shock waves driven by an ensemble of massive stars may trigger additional star formation and/or sweep the interstellar gas out of the region actively forming stars (Tenorio-Tagle \\& Bodenheimer 1988). The gas flows create a turbulent pressure which helps support the weight of the ISM (e.g. McKee 1990) and cavities which apparently enhance the distance ionizing radiation propagates (Hunter \\& Gallagher 1997; Martin 1997). This feedback from star formation may have a particularly strong influence on the evolution of low mass galaxies. Owing to their low escape velocity, Larson (1974) suggested that the loss of supernova-heated gas would begin earlier and carry away a larger fraction of their initial mass. This idea was further developed by Dekel \\& Silk (1988) who used the supernova feedback to regulate the star formation history of the evolving dwarfs. Their starburst-driven wind models were consistent with the observed mass-metallicity and mass-radius scaling relations of dwarfs when a halo similar to those produced in cold dark matter cosmological simulations was included. Mass loss has subsequently been proposed to explain a number of peculiarities about dwarf galaxies such as their abundance patterns (Marconi,Matteucci, \\& Tosi 1994) and rapid evolution at moderate redshifts (Phillipps \\& Driver 1995; Babul \\& Rees 1992). The ejection of the ISM may not be as easy as previously thought, however. In particular, the rupture of a supershell perpendicular to a galactic disk may vent much of the energy leaving most of the disk intact (DeYoung \\& Heckman 1994). Observations of dwarf galaxies reveal an environment conducive to the growth of large bubbles. Their rotation is typically nearly solid body, so shells are not sheared apart; and metallicities are generally sub-solar so cooling times are longer. Indeed, small bubbles permeate the star forming regions of the Magellanic Clouds, and a hierarchy of giant shells ($R < 300$~pc) and supergiant shells ($R \\ge 300$~pc) is plainly visible (Davies, Elliot, \\& Meaburn 1976; Meaburn 1980; Kennicutt \\et 1995). The formation of regions like 30~Doradus, which will evolve into a supergiant shell (Chu \\& Kennicutt 1994), may be thought of as the first step in the formation of a galactic outflow. Deep imaging of the ionized gas in other dwarfs yields a plethora of candidate structures for supergiant shells. Indeed, roughly one out of every four high-surface brightness dwarfs exhibit at least one shell and/or filaments (Hunter \\et 1993). It is not always obvious, however, which arcs and filaments will show the kinematic signature of an expanding shell (Hunter \\& Gallagher 1990). The kinematic evidence is mounting that some shells do breakthrough the ambient neutral gas. In the LMC, for example, the kinematics of many supershells are surprisingly quiescent compared to the giant shells (Hunter 1994). Some of these supershells are believed to be the inner ionized surface of cylindrical HI holes (Meaburn 1979; Meaburn 1980; Hunter 1994), which may have formed as a superbubble blew out perpendicular to the galactic plane. In another Magellanic irregular galaxy, \\n4449, the very large HI hole may be associated with a shell that expanded out of the galactic plane (Hunter \\& Gallagher 1997). In less luminous galaxies like the blue compact dwarf \\n1705, the expansion of the shell around the central starburst is decidedly non-spherical (Meurer \\et 1992). The kiloparsec scale, expanding shells in amorphous dwarfs (Marlowe \\et 1995) and the faint galaxy IZw18 (Martin 1996) also seem to be elongated in the general direction of the HI minor axis. At issue, however, is whether any of these disk outflows develop into freely flowing winds in which the gas actually escapes from the gravitational potential of the galaxy. Only one member of Marlowe's sample, \\n3955, was a strong wind candidate. The most convincing arguments for actual mass ejection are based on the detection of \\x emitting gas well above the galactic plane of \\n1569 (Heckman \\et 1995). The association of a hot bubble with the cavity formed by the expanding network of extended \\Ha filaments is reminiscent of the minor axis outflow from M82 (Bland \\& Tully 1988; Strickland \\et 1996; Shopbell \\et 1997), although it is not yet clear whether the dynamics of these two classes of galactic outflows are completely analogous. A more extensive kinematic census is desired to assess the frequency of blowout and the amount of mass loss. This paper presents a catalog of large-scale expanding structures in 14 nearby dwarfs. Although M82 does not strictly meet the sample selection criteria, it was added to the sample to provide a common galaxy between this study and studies of superwinds from more luminous starbursts (Heckman, Armus, \\& Miley 1990, hereafter HAM). Galaxies were selected from a volume of radius $d \\le 10$~Mpc, right ascension $4 h \\le \\alpha \\le 14~h $, and declination $\\delta \\ge -35$\\deg. An effort was made to pick the galaxies with the most intense star formation over a range in absolute luminosity from $M_B = -13.5$ to $M_B = -18.5$. Each radial velocity field was sampled with deep, high-resolution spectra of the \\Ha emission. Additional properties of the galaxies are summarized in Table~\\ref{tab:sam}. While similar scale shells are found throughout the sample, the net impact on the host galaxy's evolution may be quite varied. Two factors which largely determine the bubble's fate -- i.e. the distribution of the HI and the gravitational potential -- are not at all uniform across the sample. Hence, the prospects for mass ejection are discussed on a galaxy by galaxy basis. The results have interesting applications for the chemical evolution of dwarf galaxies and the regulation of their star formation rate. This paper is organized as follows. The observations and data reduction are described in \\S~\\ref{sec:obs}. Section~\\ref{sec:results} describes the kinematics of the ionized gas, and \\S~\\ref{sec:dynamics} discusses the dynamics of the expanding shells. Rotation curves are sketched in \\S~\\ref{sec:mdot}, and the shell expansion speeds are compared to the escape velocity. Section~\\ref{sec:sum} summarizes the main results. ", "conclusions": "\\label{sec:sum} An extensive set of \\Ha echellograms and images were used to reconstruct the large-scale kinematics of the ionized gas in 14 dwarf galaxies and M82. Details of the results for individual galaxies are included in their respective subsections of the paper, and \\fig~\\ref{fig:weaver} provides a concise summary of the shell expansion speeds and sizes. The main results regarding the formation of winds in dwarf galaxies are summarized here. \\begin{itemize} \\item The formation of supershells must be a common byproduct of massive star formation in dwarf galaxies. Expanding, supergiant ($R > 300$~pc) shells were found in all but two of the galaxies. This sample was drawn from the population of nearby dwarf galaxies with prominent arcs and/or extended filaments in their \\Ha emission, and roughly one out of four catalogued dwarf galaxies fits this description (Hunter, Hawley, \\& Gallagher 1993). Indeed, the hierarchical growth of these structures probably began in star forming regions akin to 30~Doradus in the Large Magellanic Cloud (Chu \\& Kennicut 1994). The most powerful outflows, i.e. \\n1569 and M82, were found to be composed of multiple {\\em cells} whose walls are probably the interface between polar shells. Star formation in the lowest luminosity galaxies, e.g. IZw18, also generates kiloparsec-scale shells. \\item Although many of the expanding complexes survive for 10~Myr, none older than $\\sim 20$~Myr were identified. The lack of shells older than this likely reflects their disruption timescale and provides an indirect measure of the scale height of the ISM. Alternatively, the ionization rate of the shells might drop abruptly on this timescale due to changes in the birthrate of massive stars and/or the illumination geometry. Although bright, extraplanar HI shells have not been detected in any of the galaxies in this sample, some HI holes in the LMC (Meaburn 1980) and \\n4449 (Hunter \\& Gallagher 1997) are probably relics of expanding supershells. The power requirements of the ionized supershells typically exceed the critical power for supersonic disk breakthrough, so a disruption scenario must be favored for them. The sequence of echellograms stepped across the southern lobe of \\n1569 constrains their deceleration and shows multiple velocity components at least up to 640~pc above the galactic plane. Future observational work must aim to detect the remains of the hot gas and ruptured shell following blowout. One might speculate that the quiescent filaments in \\n5253 could be fragments of a ruptured shell or clumps of infalling material ejected in a previous wind epoch. A better census of the local dwarf population would also be helpful for constraining the duty cycle of the winds. \\item Presuming the shells do rupture, the escape of hot, X-ray emitting gas from their interiors seems certain. A diffuse, thermal component of the \\x emission has been resolved in three of the galaxies in the sample, but it is only a significant fraction ($> 10\\%$) of the interstellar HI mass in M82. In contrast, much of the interstellar gas swept into the warm ionized shells probably remains bound to the galaxy. The structure of the dark matter halo has been measured in several low surface brightness dwarf galaxies with large HI disks (e.g. Carignan \\& Beaulieu 1989; Meurer \\et 1994) and appears to have a universal structure (Burkert 1995; Navarro, Eke, \\& Frenk 1996). Hence, a conservative approach to mass loss is to assume that the bursting dwarfs are embedded in similar halos. These dark halo models often do not provide enough mass to explain the HI rotation speed in the inner galaxy, however. Stars and atomic gas can account for essentially all the dynamical mass inside $R(v_{circ})$ in some galaxies like \\n1569, but the fraction varies enormously among the sample members. Although little CO emission is detected from the dwarf galaxies (e.g. Young \\et 1996), the large uncertainty in the $H_2$ to CO conversion factor does allow substantial mass contributions from molecular gas (e.g. Maloney \\& Black 1988). A dominant disk in the inner few kiloparsecs has two immediate implications. First, the disk mass contributes significantly to the gravitational acceleration of the kiloparsec-scale shells/disk outflows. Second, the observed turnover in the rotation curve may not be revealing much about the core radius of the halo -- a critical parameter for estimating the escape velocity. The maximum circular velocities in the galaxies do generally appear to be comparable to the expansion velocities of the supershells, so the escape velocities are greater than the projected shell speed. The expansion speeds along the minor axis of \\n1569 do reach values several times the maximum rotation speed. \\item The warm shells alone lift gas out of the disk at rates comparable to or greater than the current galactic star formation rates. The shells transport $10^5$ to $10^6$\\msun\\ of gas over kiloparsec-scale shells in 10~Myr and leave the sound speed high in a large volume of the ISM. This hydrodynamic mixing will be faster than a diffusion process, so the bubbles will clearly alter the chemical evolution of these galaxies. The timescales for blowout are shorter than the evolutionary timescales of most models for Type~Ia supernova progenitors, so the mass loss begins before much of the iron from the burst has been mixed into the ISM. The composition of the ejected material will depend on the duration of the wind and the composition of the ambient ISM, so their impact on the galactic chemical evolution is interwoven with the galactic star formation history. \\item Although the current kinetic energy in the large expanding structures is only comparable to the binding energy of the ISM in \\n1569, bubble blowout may still extinguish star formation in particular regions of the other galaxies. If the hot spots percolate across the dwarf irregular galaxies for many rotational periods, then a substantial fraction of the interstellar gas may still be cycled through a halo and/or lost from the galaxy. If the present mass loss rates could be sustained for 6 orbital timescales, for example, then most of the interstellar HI could be removed from the disks of six of the 15 galaxies. This global gas-dynamical feedback will be discussed in the context of the galactic star formation history in a forthcoming paper. \\end{itemize}" }, "9804/astro-ph9804023_arXiv.txt": { "abstract": "Annihilation of high energy, $\\sim 10^{21}$eV, neutrinos on big bang relic neutrinos of $\\sim 1$eV mass, clustered in the Galactic halo or in a nearby galaxy cluster halo, has been suggested to generate, through hadronic $Z$ decay, high energy nucleons and photons which may account for the detected flux of $>10^{20}$eV cosmic-rays. We show that the flux of high energy nucleons and photons produced by this process is dominated by annihilation on the uniform, non-clustered, neutrino background, and that the energy generation rate of $\\sim 10^{21}$eV neutrinos required to account for the detected flux of $>10^{20}$eV particles is $>10^{48}{\\rm erg/ Mpc}^3{\\rm yr}$. This energy generation rate, comparable to the total luminosity of the universe, is $\\sim4$ orders of magnitude larger than the rate of production of high energy nucleons required to account for the flux of $>10^{19}$eV cosmic-rays. Thus, in order for neutrino annihilation to contribute significantly to the detected flux of $>10^{20}$eV cosmic-rays, the existence of a new class of high-energy neutrino sources, likely unrelated to the sources of $>10^{19}$eV cosmic-rays, must be invoked. ", "introduction": "The Fly's Eye \\cite{Fly} and AGASA \\cite{AGASA} experiments confirmed the existence of a break in the energy spectrum of high energy cosmic rays at $\\sim5\\times10^{18}{\\rm eV}$, for which evidence existed with weaker statistics in the data of previous experiments (Haverah Park, Yakutsk, Sugar, see e.g. \\cite{Watson} for a review). Fly's Eye data also strengthen the evidence for a change in primary composition from predominantly heavy nuclei below the break to predominantly light nuclei above the break. These features strongly suggest, when coupled with the lack of anisotropy that would be expected for cosmic-rays (CRs) of Galactic origin, that below $\\sim 10^{19}{\\rm eV}$ the CRs are mostly heavy ions of Galactic origin, and that an extra-Galactic component of protons dominates above $\\sim10^{19}{\\rm eV}$. This conclusion is further supported by the fact that the CR energy spectrum is consistent with a cosmological distribution of sources of protons, with injection spectrum $dN/dE\\propto E^{-2.2}$ typically expected for Fermi acceleration \\cite{cosmology}. In particular, there is evidence for the existence of a Greisen-Zatsepin-Kuzmin (GZK) ``cutoff'', i.e. for the suppression of CR flux above $\\sim5\\times10^{19}{\\rm eV}$ expected due to interaction of protons with the microwave background radiation \\cite{GZK}. The evidence for GZK suppression is strengthened by recent AGASA data \\cite{AGASA1}. In Fig. 1, the CR spectrum reported by the Fly's Eye and the AGASA experiments \\cite{Fly,AGASA1} is compared with the flux expected for a homogeneous cosmological distribution of sources, each generating a power law differential spectrum of high energy protons $dN/dE\\propto E^{-2.2}$ (For the model calculation we have used a flat universe with zero cosmological constant, $H_0=75{\\rm km}\\ {\\rm s}^{-1}$, and time independent energy generation rate per comoving unit volume $5\\times10^{44}{\\rm erg/Mpc}^3{\\rm yr}$; The spectrum is insensitive to the cosmological parameters and to source evolution, since most of the cosmic rays arrive from distances $<1$Gpc \\cite{cosmology}). The deficit in the number of events detected above $5\\times10^{19}{\\rm eV}$, compared to a power-law extrapolation of the flux at lower energy, is consistent with that expected due to a cosmological GZK suppression. However, with current data the ``cutoff'' is detected with only $2\\sigma$ significance \\cite{Tokyo}. The number of events detected above $10^{20}$eV is consistent with that expected based on the cosmological model presented in Fig. 1 (There is an apparent ``gap'' between the highest and second highest energy events detected by the Fly's Eye \\cite{gap}. However, assuming that the cosmological model is valid, the probability that such an apparent ``gap'' would be observed is $\\sim15\\%$ \\cite{cosmology,gap}). Nevertheless, the detection of $>10^{20}$eV events does pose challenges to most models of CR production. The high energies rule out most of the acceleration mechanisms so far discussed \\cite{Hillas}, and since the distance traveled by such particles must be smaller than $100{\\rm Mpc}$ \\cite{dist} due to their interaction with the micro-wave background, their arrival directions are inconsistent with the position of astrophysical objects, e.g. jets of powerful radio galaxies \\cite{Biermann}, that are likely to produce high energy particles \\cite{obj}. Cosmological $\\gamma$-ray bursts (GRBs) are likely sources of high-energy CRs, which may account for the CR flux above $10^{19}$eV as well as for the $>10^{20}$eV events \\cite{GRBs}. This model recently gained support form GRB afterglow observations \\cite{AG}. Other models for the production of ultra-high energy CRs were suggested, where the highest energy events are produced by the decay of super-massive elementary particles related to grand unified theories (see, e.g., \\cite{Berezinsky} for recent review). Sources of such particles may be topological defects, left over from a phase transition associated with the symmetry breaking of the grand unified theory \\cite{TD}. While no firm prediction exists of the CR flux in these theories, a generic feature of the super-massive particle decay scenarios is that the injection spectrum is much harder than expected for Fermi acceleration. Therefore, this scenario can account only for the flux of $>10^{20}$eV particles, and can not simultaneously explain the origin of $10^{19}$--$10^{20}$eV CRs. It has recently been suggested that annihilation of high energy, $\\sim 10^{21}$eV, neutrinos on big bang relic neutrinos of $\\sim 1$eV mass, clustered in the Galactic halo or in a nearby galaxy cluster halo, may generate high energy nucleons and photons which may account for the detected flux of $>10^{20}$eV cosmic-rays \\cite{Weiler}. The existence of $>10^{21}$eV neutrino flux was argued plausible based on the argument that the mechanism producing the observed high energy, $>10^{19}$eV, particles, most likely protons, also produces charged pions of comparable energy, which subsequently decay to produce neutrinos. It was suggested that the generation spectrum extends well beyond $10^{20}$eV, and that while nucleons produced by a distant source, e.g. a powerful radio galaxy, lose their energy interacting with the micro-wave background, high-energy neutrinos propagate without energy losses and may annihilate on relic neutrinos, producing $>10^{20}$eV nucleons and photons at small distances that would allow them to propagate to Earth. In Sec. 2 we derive the energy density of high energy neutrinos required to account for the observed rate of $>10^{20}$eV air showers. The implications of our results are discussed in Sec. 3. ", "conclusions": "" }, "9804/hep-ph9804336_arXiv.txt": { "abstract": "The broken-symmetry electroweak vacuum is destabilized in the presence of a magnetic field stronger than a critical value. Such magnetic field may be generated in the phase transition and restore the symmetry inside the bubbles. A numerical calculation indicates that the first-order phase transition is delayed but may be completed for a sufficient low value of the Higgs mass unless the magnetic field is extremely high. ", "introduction": "It has been found that very strong magnetic fields are capable of destabilizing the electroweak vacuum by forming a vector boson $W^{+}W^{-}$ condensate and restoring the symmetry \\cite{amb}. The required field can only be thought to have existed at the very beginning of the universe and one of the possibilities is that it was generated during the electroweak phase transition \\cite{baym,cheng}. This primordial field may have been subsequently the seed of the present galactic magnetic field \\cite{enq}. One may wonder whether the restoration of symmetry caused by this strong magnetic field can delay the electroweak phase transition. In particular, if it is of first-order the magnetic field might avoid its completion through the bubble mechanism. The simplest way to see why a strong magnetic field can destabilize the electroweak vacuum is to consider the energy of a charged spin-one particle interacting with a uniform magnetic field along the 3-axis \\begin{equation} E_{N}^{2}=p_{3}^{2}+m_{0}^{2}+\\left( 2N+1\\right) eB-geB\\quad . \\end{equation} For the lowest Landau level $N=0$ if the gyromagnetic factor $g$ is $2$ as occurs in the $W$ case, it is clear that the effective mass will become zero for \\begin{equation} B_{c}=\\frac{m_{W}^{2}}{e}\\simeq 10^{24}G\\quad . \\label{eq2} \\end{equation} This expression is analogous to that of the critical electric field required to create pairs through tunneling. If one wishes to calculate the decay probability of the vacuum, one must evaluate \\begin{equation} Z=<0|e^{-iHt}|0>=e^{-it\\left( E_{vac}-i\\frac{\\Gamma }{2}\\right) }\\quad . \\end{equation} In Euclidean metric the one-loop amplitude for a scalar field depends on $ \\det \\left( -D_{E}^{2}+m^{2}\\right) $ with $D_{E\\mu }=\\partial _{\\mu }-ieA_{\\mu }$, being $D_{E4}=iD_{0}$. Using the Schwinger proper time method \\cite{schw} one obtains \\begin{equation} \\ln Z=\\int_{0}^{\\infty }\\frac{ds}{s}tre^{-\\left( -D_{E}^{2}+m^{2}\\right) s}\\quad . \\label{e2} \\end{equation} For constant electromagnetic fields the trace is known to give the vacuum energy density \\cite{schm} \\begin{equation} \\rho =-\\int_{0}^{\\infty }\\frac{ds}{s}\\frac{e^{-m^{2}s}}{\\left( 4\\pi s\\right) ^{2}}\\left[ \\frac{es\\sqrt{E^{2}-B^{2}}}{\\sin \\left( es\\sqrt{E^{2}-B^{2}} \\right) }-1\\right] \\quad , \\label{e1} \\end{equation} where the $-1$ comes from subtracting $\\rho \\left( A=0\\right) $. This integral has a logarithmic divergence for $s=0$ which can be absorbed renormalizing fields and charge \\cite{schw}. In Eq.(\\ref{e1}) for $E>B$ the integral has poles in the $s$-axis which give origin to an imaginary part corresponding to pair creation. We will be instead interested in the case of $E=0$ and constant magnetic field for which \\begin{equation} \\rho =-\\int_{0}^{\\infty }\\frac{ds}{s}\\frac{e^{-m^{2}s}}{\\left( 4\\pi s\\right) ^{2}}\\left[ \\frac{esB}{\\sinh \\left( esB\\right) }-1\\right] \\end{equation} that has no poles. For the spin-1 $W_{\\mu }$ case we adopt the view that the only modification to $\\ln Z$ is the interaction of spin with magnetic field $2e\\mathbf{B}\\cdot \\mathbf{s}$ in the exponent of Eq.(\\ref{e2}). Now the trace must be performed on momentum and spin states where the latter involves this added interaction to give \\begin{equation} \\ln Z=\\int_{0}^{\\infty }\\frac{ds}{s}\\left( e^{-2eBs}+e^{2eBs}+1\\right) tre^{-\\left( -D_{E}^{2}+m_{W}^{2}\\right) s}\\quad . \\end{equation} Since the remaining trace is equal to the scalar case, the relevant part of the vacuum energy density is \\begin{equation} \\rho =-\\int_{0}^{\\infty }\\frac{ds}{s}\\frac{e^{-m_{W}^{2}s}}{\\left( 4\\pi s\\right) ^{2}}\\left[ \\frac{esB}{\\sinh \\left( esB\\right) }2\\cosh \\left( 2eBs\\right) \\right] \\quad . \\end{equation} This expression has no poles but diverges for $s\\rightarrow \\infty $ when $ B>B_{c}=m_{W}^{2}/e$ due to the gyromagnetic factor $2$ of the $W$ boson, which would not occur either for $g=1$ or for the $s=1/2$ case. This divergence is an indication of the vacuum instability for large magnetic field. The decay rate should be evaluated in the more realistic situation of $B$ increasing with time, with the consequent generation of an electric field. In our calculation of the next section we will not take into account the evolution with time of the magnetic field but we will consider that when its value is larger than the critical one in the region of a bubble containing the broken-symmetry vacuum, the bubble will be destroyed. ", "conclusions": "We have seen that the highest possible magnetic field together with the most favorable law for having homogeneous field in regions of increasing size might have cosmological consequences through the non-completion of the first-order electroweak transition through a bubble mechanism. Therefore the usual electroweak baryogenesis due to bubble expansion would be affected. But one must notice that also the homogeneous increase of $\\varphi $ can produce a matter-antimatter asymmetry. This is because there will be a baryonic chemical potential related to the time variation of the $CP$ violating phase $\\theta $. The resulting baryonic density will depend on the variation $\\Delta \\theta $ in the interval when the sphalerons are active due to the smallness of $\\varphi $. For a weakly first-order transition, an advantage of this mechanism compared to the bubble one is that the baryonic density would not be erased in the broken phase because here the value of $ \\varphi $ is larger due to the delay of the phase transition. One may remind that this problem is also avoided by the baryogenesis in cosmic strings but paying the price of a suppression factor in the active volume. However, it is unlikely that such a strong and large size primordial magnetic field has occurred, and for more acceptable fields the effect would be only a small decrease of the temperature for the completion of a first-order transition. We have studied the influence of the magnetic field on the phase transition using the easiest model, i.e. the standard model and not the MSSM where presumably the first-order phase transition can occur for not too light Higgs mass \\cite{care}. The fact that we obtain the completion of the first-order transition without magnetic field for $m_{H}\\simeq 70GeV$, not far below the experimental bound is probably due to the definition that it occurs when the bubbles touch each other without taking into account their scattering. But we believe that the general conclusions on the magnitude of the effect do not depend on the details of the used electroweak model. Regarding further developments of this calculation, it would be important to evaluate the vacuum decay rate caused by a time dependent magnetic field in order to consider more carefully its effect on the bubbles instead of taking the simplification of assuming their disappearance as soon as the magnetic field is larger than the critical value. \\strut \\strut \\strut \\textbf" }, "9804/astro-ph9804090_arXiv.txt": { "abstract": "We have computed finite temperature corrections to the electron-hadron scattering cross sections. These are based upon the renormalized electron mass and the modified density of states due to the presence of a background thermal bath. It is found that the electron-hadron thermal transport scattering cross section can be much larger than the zero temperature one. In the case of electron-neutron transport scattering, we find $\\sigma_{ne}(T) / \\sigma_{ne} (T=0) \\simeq 5$ at $T \\simeq 0.1 \\, MeV$. \\vskip 0.3cm ", "introduction": "Finite temperature effects on elementary processes are significant from the point of view of cosmology and astrophysics. The early universe is usually described as a hot gas of particles in nearly thermodynamical equilibrium. Temperature effects enter through the statistical distribution functions. These can renormalize the masses and the wave functions. These renormalized masses and wave functions can then affect scattering processes and decay rates. Several authors \\cite{donoghue} have generalized the electron-mass and wave-function renormalization to all temperatures and densities. Dicus et al. \\cite{dicus} and independently, Cambier et al. \\cite{cps} included the finite temperature effects on weak reaction rates in calculations of standard big-bang nucleosynthesis $(BBN)$. They obtained the corrected light-element abundance and found that the corrections are only of order of a few percent. After that, Saleem \\cite{saleem} included the effects of the electron mass shift at finite temperature on $BBN$ and Baier et al. \\cite{baier} examined the finite temperature radiative corrections to the weak neutron-proton decay rates. More recently, Fornengo et al. \\cite{jwkim} have considered the finite temperature effects on the neutrino decoupling temperature which is important in the evolution of the early universe. In the present work we consider finite temperature corrections to electron-hadron scattering which is important for baryon inhomogeneous cosmologies. Baryon inhomogeneities might have been produced during the cosmological quark-hadron phase transition in the early universe \\cite{witten}. If such inhomogeneities were present, then the different diffusion lengths for neutrons and protons could lead to the formation of high-baryon density proton-rich regions and low-baryon density neutron-rich regions. The light element nucleosynthesis yields from such regions can differ significantly from those of standard homogeneous big-bang nucleosynthesis \\cite{AHS}. In view of the importance of using light-element yields from the $BBN$ to constrain the baryon-to-photon ratio as well as various cosmological and particle physics theories, such inhomogeneous models must be examined seriously. It is therefore important to quantify the effects of baryon diffusion as accurately as possible. In this regard Applegate, Hogan and Scherrer (AHS) \\cite{AHS} have calculated the diffusion rate of baryons through the electron-positron plasma in the early universe. Subsequently, several authors used their results in calculations of inhomogeneous $BBN$ \\cite{applegate,mathews}. In $AHS$ it was suggested that the diffusion coefficients could be derived from the mobility of the heavy particles, and that the mobility is determined from the distribution functions of the background plasma and the transport cross section. It is important, therefore, to carefully quantify the values of the distribution functions and the transport cross sections. However, in all previous baryon diffusion coefficient calculations, vacuum transport scattering cross sections have been used. Therefore, in order to estimate the baryon diffusion coefficients more precisely, in the present work we take into account the finite temperature effects in the calculation of baryon diffusion coefficients at temperatures $\\lsim \\; MeV$. Specifically, we calculate the transport scattering cross section of elastic electron-hadron scattering at finite temperature. Here we shall treat hadrons as particles which have an internal structure and an anomalous magnetic moment (although we do not have a good field theory for the magnetic moments of protons or neutrons at finite temperature). Also, we assume that their internal structure is not affected by finite temperature since their mass is more than about 1000 times the temperatures of interest. The plan of the paper is as follows. In Section \\ref{sec:finite}, we discuss how to include finite temperature effects in the calculation. In particular, we will briefly discuss the effective mass of an electron in the MeV temperature range. In Section \\ref{sec:scattering}, we evaluate the electron-hadron transport scattering cross section at finite temperature. Finally, we summarize our results and discuss some astrophysical applications. We shall employ units in which $\\hbar = k_B = c = 1$, except when specific units must be attached to a result. ", "conclusions": "We have calculated temperature-dependent electron-hadron transport cross sections. These are important, for example, in the calculation of baryon diffusion coefficients at finite temperature. The major motivation here has been to investigate whether finite temperature effects can significantly change the baryon transport cross section $\\sigma_t$. In this work, we have treated hadrons as particles which have an internal structure and an anomalous magnetic moment. Also, we have assumed that their internal structure is not affected by the finite temperature since their mass is more than about 1000 times the temperature of interest. Two major features of the finite temperature effects on the light particles have been included in the calculation: (1) finite temperature Dirac spinors which are recast into the form of an effective electron mass ; (2) finite temperature modifications to the phase space distribution of the electrons. We find that, for $m_0 < T$, both $\\sigma_{ne}(T)$ and $\\sigma_{pe}(T)$ approach the ultrarelativistic limit (where the electron mass can be ignored). In the case of electron-proton scattering, we have compared it with the Coulomb scattering cross section at finite temperature. In particular, for the case of electron-neutron transport scattering, we find $\\sigma_{ne}(T) / \\sigma_{ne} (T=0) \\simeq 5$ at $T \\simeq 0.1 \\, MeV$. In conclusion, the baryon diffusion coefficients which affect baryon inhomogeneities during big-bang nucleosynthesis could be changed significantly by our temperature dependent electron-hadron transport cross sections. Up to the time of weak decoupling ($T \\simeq 1 \\, MeV$) there is little change in the cross sections. However, during the epoch of nucleosynthesis ($T \\lsim 0.2 MeV$) when baryon diffusion is most important, the transport cross sections increase as the temperature decreases. On the other hand, baryon diffusion at low temperature is strongly affected by proton-neutron scattering for which these finite temperature effects are insignificant. Clearly, a study of the effects of these new cross sections on the baryon diffusion coefficients and inhomogeneous primordial nucleosynthesis is desired. These will be the subject of a subsequent paper. \\vspace{0.5cm} {\\bf Acknowledgements.} \\noindent The author (ISS) would like to thank P. Marronetti and Prof. S. Rhie for their helpful comments. ISS also acknowledges the support by the Korea Research Foundation (KRF) for the Post-Doctoral Fellowship at University of Notre Dame. This work supported in part by DOE Nuclear Theory Grant DE-FG02-95ER40934." }, "9804/astro-ph9804059_arXiv.txt": { "abstract": "To determine the magnification of an extended source caused by gravitational lensing one has to perform a two-dimensional integral over point-source magnifications in general. Since the point-source magnification jumps to an infinite value on caustics, special care is required. For a uniformly bright source, it has been shown earlier that the calculation simplifies if one determines the magnification from the area of the images of the extended source by applying Green's theorem so that one ends up with a one-dimensional integration over the image boundaries. This approach is discussed here in detail, and it is shown that it can be used to yield a robust and efficient method also for limb-darkened sources. It is also shown that the centroid shift can be calculated in a similar way. ", "introduction": "For fitting light curves for the ongoing microlensing events, there is a need for robust and efficient methods for calculating the magnification of extended sources, which are not limited to point-lenses. Among the observed events, the presence of binary lenses is a reality (Dominik \\& Hirshfeld~\\cite{MLMC1let},\\cite{MLMC1}; Udalski et al.~\\cite{OGLE7}; Alard et al.~\\cite{DUO2}; Bennett et al.~\\cite{MLMC9}), and planetary events involve a special case of a binary lens. In addition, for some configurations, the light curve for a limb-darkened source will differ significantly from that of a uniformly bright source. The limb-darkening effect has recently been observed in the galactic microlensing event MACHO 97-BLG-28, which involves both an extended source and a binary lens, by the PLANET collaboration (Albrow et al.~1998a,b\\nocite{M28Planet1}\\nocite{M28Planet2}); the fitting has been done by myself using the algorithm described in this letter. If one wants to integrate the point-source magnification in two dimensions one has to take special care of the position of the caustics, where the point-source magnification becomes infinite. While this integration can be performed easily for a point-mass lens (e.g. Schneider et al.~\\cite{SEF}, p. 313; Witt \\& Mao~\\cite{WM}; Sahu~\\cite{Sahu}; Dominik~\\cite{DoDiss}), this would be a difficult task for a general lens (e.g. a binary lens), especially at a cusp singularity. In contrast, the area of the images of the extended source and therefore its magnification remains continuous when the source hits a caustic. The determination of the extended source magnification from the boundaries of the image areas has been used to analyze the images of background galaxies behind a cluster of galaxies (Dominik~\\cite{DoDipl}). The image boundaries can be obtained with a contour plot of an implicit function describing the source boundary in the lens plane (Schramm \\& Kayser~\\cite{SK}). This method has been expanded with routines for correcting, testing and finally analyzing the contour line in order to produce an efficient and safe algorithm (Dominik~\\cite{DoAstro}). In that paper, it is noted that it is easy to analyze the images from the contour line data, and an example is given, where quantities such as the area, width, length and curvature of the image have been determined. Concerning microlensing light curves, it has been noted by Bennett \\& Rhie (\\cite{BenRhie}) that it is advantageous to integrate in the lens plane rather than in the source plane to determine the magnification of an extended source. For uniformly bright sources, Gould \\& Gaucherel (\\cite{GouGau}) proposed applying Green's theorem so that only one integration along the image boundary must be performed rather than two over the image area. This approach is identical to that used earlier (Dominik~\\cite{DoDipl},~\\cite{DoAstro}). The contour plot method is the most convenient way to obtain data points on the image boundary from which the area can be calculated. In Sect.~2, this general approach is described, Sect.~3 gives details for a uniformly bright source, and Sect.~4 shows how this approach can also be used for limb-darkened sources, in which case an easy-to-perform two-dimensional integration remains. In Sect.~5, the calculation of the centroid shift is discussed. ", "conclusions": "" }, "9804/astro-ph9804329_arXiv.txt": { "abstract": "s{ Knowledge of the constants that describe the current cosmological world model, $H_0$, $t_0$, and the three $\\Omega$s is central to physical cosmology. Although there is a vast range of suggested tests and existing constraints much of the recent discussion of cosmological constants involves three time variable photospheres: Cepheids, SNIa, and the CMB last scattering surface. These concluding remarks for the Moriond XXXIII meeting are made at a time when many of the established methods have made careful, interesting, statements about the values of various cosmological constants based on data of small random errors. The flood of new data over the next few years will lead to a satisfying increase in the precision of both direct and model dependent estimates of the main cosmological parameters. } ", "introduction": "Commenting on the progress being made in estimating the values of the cosmological parameters has perils not unlike those of critiquing a great artwork, perhaps an opera, while it is being staged for the first time. The problem is that there is no working consensus for the values of the cosmological constants. There are rather wide compromise ranges which accommodate most of the derived values. However the values of the cosmological constants at one end of a compromise range are completely incompatible with those at the other, both in physical meaning and in stated measurement error. An aspect of cosmological parameter determination that is fascinating for any interested scientist is that it continues to reward an integrated view of the entire subject from trigonometric parallaxes to photometric response functions to radio interferometric mapping to X-ray plasma analysis to the outer limits of particle theory, to name but a few. Martin Rees has described Cosmology as the ``Grandest of the Environmental Sciences'' \\cite{rees} which serves to remind us of the difficulty of relating the observations of the universe to the simple cosmological models of interest. Although the FRW model and an interest in its parameters have been around for a long time, it was not until the 1980's when development of solid state detectors of very low noise and very high quantum efficiency allowed virtually every waveband used by astronomy to greatly increase both the quality and the abundance of data. The objects now examined range from nearby white dwarfs to galaxies at redshifts beyond 5, to precise measurements of the CMB radiation all over the sky. The following discussion discusses the broad comprise ranges for the basic cosmological parameters. In most cases the data have small random errors. The problems for all methods come in calibration and model uncertainties. For instance Cepheids distances have random errors of about 1\\%, nevertheless the total error in the Hubble constant is generally quoted to be at least 10 times larger as a result of potential calibration and systematic errors. Even those apparently large error budgets are hard won from vast efforts to control, measure and remove systematic errors. The true error ranges of these are continuing to shrink significantly and steadily, such that sometime in the next decade our understanding will undergo a phase transition, hopefully a crystallization not a meltdown. ", "conclusions": "The current situation is precisely why the measurement of the cosmological parameters is the primary activity of many astronomers and astrophysicists. The exciting likelihood is that most of the major cosmological constants will be known to a satisfying degree of on the time scale of a decade. The Moriond meetings provide an ideal format for frank discussion of extremely controversial cosmological issues. I thank the organizers for the splendid job they did in managing to bring us all together and providing a never ending flow of food and stimulus for mind and body." }, "9804/astro-ph9804144_arXiv.txt": { "abstract": "The proton burning process $p+p\\rightarrow d +e^+ +\\nu_e$, important for the stellar evolution of main-sequence stars of mass equal to or less than that of the Sun, is computed in effective field theory using chiral perturbation expansion to the next-to-next-to-leading chiral order. This represents a model-independent calculation consistent with low-energy effective theory of QCD comparable in accuracy to the radiative $np$ capture at thermal energy previously calculated by first using very accurate two-nucleon wavefunctions backed up by an effective field theory technique with a finite cut-off. The result obtained thereby is found to support within theoretical uncertainties the previous calculation of the same process by Bahcall and his co-workers. ", "introduction": "The proton fusion reaction \\be p + p \\rightarrow d + e^+ + \\nu_e \\label{pp}\\ee which plays an important role for stellar evolution and -- as the dominant neutrino source -- for the solar-neutrino problem, has quite a long history of investigation. Indeed the reaction rate of this process (hereafter called the $pp$ rate) was first calculated by Bethe and Critchfield (\\cite{bethe38}). Salpeter (\\cite{salpeter}) recalculated the $pp$ rate using the {\\em effective range approximation} and argued that the relevant nuclear matrix element squared could be estimated with an accuracy of the $\\sim$5 \\% level. (The $pp$ rate itself was subject to much larger uncertainty, $\\sim$20 \\%, because of the limited precision with which the Fermi coupling constant was known at that time.) Bahcall and May (\\cite{bahcall69}) examined the dependence of the $pp$ rate on explicit forms of the two-nucleon wavefunctions generated by two-parameter nuclear potentials of various forms adjusted so as to reproduce the scattering length and effective range (for the $pp$ channel) and the low-energy properties of the deuteron (for the $np$ channel). The $pp$ rate was found to vary by $\\sim 1.5$ \\% corresponding to the changes in the deuteron wavefunction, and by $\\sim 1.2$ \\% due to the change in the $pp$ wavefunction. The most updated work along this line was done by Kamionkowski and Bahcall (\\cite{bahcall94}) employing deuteron wavefunctions obtained from much more accurate potentials such as the Argonne $v_{14}$, $v_{18}$, Urbana $v_{14}$, super-soft-core (SSC) and Reid soft-core potentials. Changes in the $pp$ rates arising from the different potentials were found to be $\\sim 1\\ \\%$. Thus it seems that the presently available calculated $pp$ rate is robust and needs no further scrutiny, the famous solar neutrino problem remaining unresolved from this angle and hence persisting as one of the outstanding unsolved problems in astrophysics (Bahcall \\cite{unsolved}). There are however two reasons for revisiting this issue. One is that while the calculated $pp$ rate seems to have converged to a ``canonical\" value given in (Kamionkowski \\& Bahcall \\cite{bahcall94}, hereafter KB), there lingers the unsettling feeling that the strong interaction involved in nuclear physics of the two-nucleon systems is infested with uncontrollable uncertainties associated with model dependence in the treatment, making it difficult to assess the accuracy achieved. Thus it is not unexpected that this canonical value will be -- as has been in the past -- challenged. Indeed it has recently been argued by Ivanov et al. (\\cite{ivanov}) that the nonrelativistic potential models used in the previous works could be seriously in error. They show that in their version of a {\\it relativistic field theory model}, the $pp$ rate comes out to be as big as $2.9$ times the previous estimates.\\footnote{They use a procedure that seems to disagree with other physical properties of low-energy $pp$ systems, as was pointed out by Bahcall and Kamionkowski (1997). It has also been pointed out by Degl'Innocenti et al. (\\cite{Degl}) that such a large deviation from the value used by KB would be inconsistent with helioseismology in the Sun.} Should their new result turn out to be correct, it would have profound consequences on theories of stellar evolution in general and on the solar neutrino problem in particular. In a nutshell, the issue comes down to whether or not a more general framework such as relativistic field theory would invalidate the calculation made in the traditional nonrelativistic potential models. The claim of the authors in (Ivanov et al. \\cite{ivanov}) is that it indeed does. Our aim is to address this issue using a low-energy effective field theory of QCD that has found a quantitative success in other nuclear processes. The second reason is really more theoretical, independent of the above important astrophysical issue. Along with the thermal $np$ capture, the proton fusion process is the simplest nuclear process amenable to an accurate calculation -- something rare in hadronic physics -- and it is of interest on its own to test how well a calculation faithful to a ``first-principle approach\" can tackle this problem. In particular, we are interested in checking how accurately the effective field theory approach, found to be stunningly successful for the $np$ capture $n+p\\rightarrow d+\\gamma$, low-energy NN scattering and static properties of the deuteron (Park, Min, \\& Rho \\cite{pmr_PRL}; Park et al. \\cite{pkmr}), fares with the proton fusion problem, the weak interaction sector of the Standard Model. The strategy we shall adopt here is quite close to that used for the $np$ capture process (Park et al. \\cite{pmr_PRL}). We shall use chiral perturbation theory to the next-to-next-to-leading order (NNLO) in chiral counting; as defined precisely below, this corresponds to $O(Q^3)$ relative to the leading-order term. This is roughly the same order as considered for the $np$ capture. However the relative importance of terms of various chiral orders is somewhat different here. As we explain later, in the present case, the corrections to the leading order are not suppressed by what is called the ``chiral filter\" \\footnote{The chiral filter phenomenon is explained in detail in (Park, Min, \\& Rho, \\cite{pmr_Report}). Crudely stated, it refers to the general feature that whenever one soft-pion exchange is allowed by kinematics and selection rules, it should give a dominant contribution with higher-order (or shorter-range) terms strongly suppressed. A corollary to this is that whenever one soft-pion exchange is not allowed, all higher-order terms {\\it can be} important, making chiral perturbation calculation generically less powerful.} and so the accuracy with which these can be calculated is not as good as in the $np$ capture case. Even so, using the argument developed in (Park et al. \\cite{pkmr}), we shall suggest that the procedure used here provides a model-independent result in the same sense as in (Park et al. \\cite{pmr_PRL}, Park et al. \\cite{pkmr}). To streamline the presentation, we first give our result and then discuss (as briefly as possible) how we arrive at it in the rest of the paper. Apart from the meson-exchange contributions which are of order of ${\\cal O}(Q^3)$ and which for the reason mentioned above and further stressed in our concluding section, are the main uncertainty to the order considered, our chiral perturbation theory result in terms of the reduced matrix element $\\Lambda$ defined in (Bahcall \\& May \\cite{bahcall69}) is \\bea \\Lambda_{\\chi PT}^2 = (1 \\pm 0.003) \\times 6.93 \\label{ChPT} \\eea where the uncertainty is due to experimental errors.\\footnote{ Our theoretical uncertainty is, if very conservatively estimated, about $0.1\\ \\%$.} The above result is to be compared with the value obtained by Kamionkowski and Bahcall (KB) \\be \\Lambda_{\\rm KB}^2 = \\left(1 \\pm 0.002^{+0.014}_{-0.009}\\right) \\times 6.92\\, . \\label{bahcallS}\\ee As we shall explain, there are some differences in details between our calculational framework and that of KB, but our final numerical result is in good agreement with that of KB and disagrees with that of Ivanov et al. (\\cite{ivanov}). The paper is organized as follows. In Section \\ref{2}, our strategy for carrying out a chiral perturbation calculation for two-nucleon systems is outlined. Our approach here is similar to the one used in the previous calculation of the $np$ capture process. We shall sketch a justification of this approach from the standpoint of low-energy effective field theory of QCD (with concrete supporting evidence summarized in Subsection \\ref{cutoff}). Section \\ref{chiralcounting} describes chiral counting of the terms appearing in the relevant weak current. In Section \\ref{wavefunction}, the wavefunctions for the initial $pp$ state and the final $d$ state are specified. Our numerical results are given in Section \\ref{numbers}, and a brief discussion including a comment on the main uncertainty in the calculation is given in Section \\ref{discussion}. ", "conclusions": "\\label{discussion} Following the procedure of chiral perturbation theory proven to be highly successful for the thermal $np$ capture process, we have calculated the $pp$ fusion rate to $\\calO (Q^3)$ in chiral counting relative to the leading single-particle Gamow-Teller matrix element. {\\it To the order considered}, the error involved in the calculation is small, $\\lesssim 1$ \\%. This result, given a justification from a cut-off effective field theory of low-energy QCD as in the case of the $np$ capture (Park et al. \\cite{pkmr}), supports the canonical value of Bahcall et al. and does not support the ``relativistic field theory model\" result of Ivanov et al. (\\cite{ivanov}). The main caveat in this calculation is in the meson-exchange contribution which comes out to be about 4 \\% when calculated to the chiral order $\\calO (Q^3)$. At the next order, $\\calO (Q^4)$, loops and higher-order counter terms enter, so that there is no reason to believe that they are negligible compared with the $\\calO (Q^3)$ tree contributions. (For instance, it could be lowered to about $1\\sim 2$ \\% instead of $\\sim 4$ \\% found here). This aspect is different from the case of the $np$ capture where the chiral filter mechanism assures the dominance of the tree-order pion exchange-current contribution. Here the absence of the chiral filter phenomenon {\\it can} allow higher-order (loop) terms to figure equally importantly as the tree-order terms. {}From this viewpoint it is not surprising that a model calculation of the terms of $\\calO (Q^4)$ and higher based on the vector-meson exchange and form factors (Bargholtz \\cite{bar79}) indicates that there can be a considerable suppression of the tree-order correction. Although such a reduction in the exchange-current contribution seems to go in the right direction for beta decays of higher-mass nuclei (Carlson \\cite{carlson}), it is probably unsafe to import the result of such model calculations into our work. To go to $\\calO (Q^4)$ or above in our theory at which heavy mesons and form factors come in, a large number of Feynman graphs of the same chiral order have to be computed on the same footing to assure chiral symmetry, and such a calculation has not been done yet. In the absence of consistent calculations, our attitude is that we should not attach any error estimates on terms not accounted for to the chiral order computed. Calculating the higher-order terms will be left for a future exercise." }, "9804/astro-ph9804002_arXiv.txt": { "abstract": "Time-resolved eclipse spectroscopy of the nova-like variable UX~UMa obtained with the HST/FOS on 1994 August and November is analyzed with eclipse mapping techniques to produce spatially resolved spectra of its accretion disc and gas stream as a function of distance from disc centre. The inner accretion disc is characterized by a blue continuum filled with absorption bands and lines which cross over to emission with increasing disc radius, similar to that reported by Rutten et~al (1994) at optical wavelengths. The comparison of spatially resolved spectra at different azimuths reveals a significant asymmetry in the disc emission at UV wavelengths, with the disc side closest to the secondary star showing pronounced absorption by an `iron curtain' and a Balmer jump in absorption. These results suggest the existence of an absorbing ring of cold gas whose density and/or vertical scale increase with disc radius. The spectrum of the infalling gas stream is noticeably different from the disc spectrum at the same radius suggesting that gas overflows through the impact point at disc rim and continues along the stream trajectory, producing distinct emission down to $0.1\\; R_{L1}$. The spectrum of the uneclipsed light shows prominent emission lines of Ly$\\alpha$, N\\,{\\sc V} $\\lambda 1241$, Si\\,{\\sc IV} $\\lambda 1400$, C\\,{\\sc IV} $\\lambda 1550$, He\\,{\\sc II} $\\lambda 1640$, and Mg\\,{\\sc II} $\\lambda 2800$, and a UV continuum rising towards longer wavelengths. The Balmer jump appears clearly in emission indicating that the uneclipsed light has an important contribution from optically thin gas. The lines and optically thin continuum emission are most probably emitted in a vertically extended disc chromosphere + wind. The radial temperature profiles of the continuum maps are well described by a steady-state disc model in the inner and intermediate disc regions ($R \\leq 0.3 R_{L1}$). There is evidence of an increase in the mass accretion rate from August to November (from \\.{M}$= 10^{-8.3\\pm 0.1} \\;{\\rm to}\\; 10^{-8.1\\pm 0.1}\\; M_{\\odot} \\; yr^{-1}$), in accordance with the observed increase in brightness. Since the UX\\,UMa disc seems to be in a high mass accretion, high-viscosity regime in both epochs, this result suggests that the mass transfer rate of UX~UMa varies substantially ($\\simeq 50$ per cent) on time scales of a few months. It is suggested that the reason for the discrepancies between the prediction of the standard disc model and observations is not an inadequate treatment of radiative transfer in the disc atmosphere, but rather the presence of additional important sources of light in the system besides the accretion disc (e.g., optically thin continuum emission from the disc wind and possible absorption by circumstellar cool gas). ", "introduction": "Accretion discs are an important phenomenon in astrophysics, invoked to solve a wide range of astrophysical problems ranging from planetary formation to quasar energetics (Frank, King \\& Raine 1992). Although considerable effort in both observation and theory has been invested over the past decade, the structure and underlying physics of accretion discs remains poorly understood. Major unsolved problems include the nature of the viscosity mechanism -- responsible for the spiraling inward of the disc material -- (the angular momentum problem), the fate of the kinetic energy expended at the inner edge of the accretion disc (the boundary layer problem), the vertical structure of the disc (the Balmer decrement problem), and the outflow of matter in connection with a disc wind (possibly a solution to, or at least an element of, the boundary layer problem). Progresses in solving these issues has been hampered because most of the previous observational constraints provided only the spectrum of the total light from the disc. A better understanding of the physics of accretion discs requires spatially-resolved studies. Cataclysmic Variables (CVs) are mass-exchanging binary systems containing a white dwarf and a late-type star (Warner 1995). If the white dwarf is not strongly magnetized ($B < 10^{6}$~G) an accretion disc is formed. Accretion discs in non-magnetic CVs cover a range of accretion rates and viscosity states. For example, {\\em dwarf novae} undergo large outbursts ($\\Delta m = 3-5$~mag, typical duration of 5-10 days) which reflects changes in the structure of the discs -- from a cool, optically thin, low viscosity state to a hot, optically thick, high viscosity state -- and which are usually parameterized as a large change in the mass accretion rate ( \\.{M}$= 10^{-11} \\; M_\\odot \\; yr^{-1} \\mapsto 10^{-9} \\; M_\\odot \\; yr^{-1}$. See, e.g. Pringle, Verbunt \\& Wade 1986). On the other hand, {\\em nova-like} variables seem to be permanently in a high viscosity state, presumably as a result of the fact that the accretion rate is always high. Because the nature of the other constituents in these systems -- the white dwarf and the normal star -- are reasonably well understood, and because orbital variations often provide considerable insight into the system geometry, non-magnetic CVs are the ideal laboratories for understanding accretion discs. Eclipsing systems are particularly useful since the occultation of the accretion disc by the late-type star provides information about the disc's spatial structure through the eclipse shape. The eclipse mapping method (Horne 1985, 1993; Rutten, van Paradijs \\& Tinbergen 1992; Baptista \\& Steiner 1993) assembles the information contained in the eclipse shape into a map of the disc surface brightness distribution. When applied to time-resolved spectroscopy through eclipses this technique delivers the spectrum of the disc at any position on its surface. Information on the radial dependence of the temperature and vertical temperature gradients (for optically thick regions), or temperature, surface density and optical depth (where the disc is optically thin) can be obtained by comparing such spectra with the predictions of models of the vertical disc structure. The spatial structure of the emission-line regions over the disc can be similarly mapped from data of high spectral resolution. Furthermore, by studying the time-variations in the structure of accretion discs of dwarf novae undergoing outbursts it may be possible to uncover the nature of the (so far unknown) viscosity mechanism which drives accretion discs. UX~UMa is a well known, bright ($V \\simeq 12.5$) eclipsing nova-like variable with an orbital period of 4.72~hr. Eclipse mapping in broad-bands (Horne 1983; Rutten et~al. 1992) shows that its accretion disc is optically thick and is close to a steady state at a mass accretion rate of $\\simeq 10^{-8} \\; M_\\odot \\; yr^{-1}$. The broad-band mapping was extended to spectrally-resolved mapping in the optical range by Rutten et al. (1993, 1994). Their results show that the continuum becomes fainter and redder with disc radius -- reflecting a radial temperature gradient -- and reveal that the Balmer lines are seen in absorption in the inner disc but in emission in the outer disc. Baptista et al. (1995) performed a similar study using HST data in narrow spectral windows about the C\\,{\\sc IV} 1550 and He\\,{\\sc II} 1640 line regions. This study showed that the UV continuum reasonably follows the $T \\propto R^{-3/4}$ law for steady mass accretion, confirming the results from the optical analysis. The C\\,{\\sc IV} and He\\,{\\sc II} line profiles are dominated by emission from the disc wind. Spatially-resolved spectra reveal that these lines appear as narrow absorption features at disc centre and change with increasing radius to broad emission in the outer disc regions besides showing large uneclipsed components. This behaviour is similar to that found for the Balmer lines and suggests that these optical lines may also have a wind component. In this paper, we report on the ultraviolet (UV) and optical mapping of the accretion disc and gas stream of UX~UMa, based on observations made with the {\\it Faint Object Spectrograph} (FOS) on the {\\it Hubble Space Telescope} (HST). The reader is referred to Knigge et~al. (1998a) for an initial description of these observations, with emphasis on the spectral properties of the integrated spectra of the accretion disc, the bright spot and the uneclipsed light. Sect.\\,\\ref{observa} describes the data and its reduction. The extraction of narrow-band light curves and their analysis with eclipse mapping techniques are described in Sect.\\,\\ref{analise}. Sect.\\,\\ref{results} presents and discusses spatially resolved spectra of the accretion disc and the gas stream region as a function of distance from disc centre, the spectrum of the uneclipsed light, and the radial temperature distribution in the ultraviolet. The possible influence of the assumed eclipse geometry on the results is addressed in Sect.\\,\\ref{geo}. Sect.\\,\\ref{discuss} discusses the implications of the results in the context of disc atmosphere models. The results are summarized in Sect.\\,\\ref{conclusao}. ", "conclusions": "\\label{discuss} Knigge et~al. (1998a) show that disc models constructed as ensembles of stellar atmospheres provide poor descriptions of the observed integrated spectrum of UX~UMa. The disc model spectra are too blue at ultraviolet wavelengths and overpredict the magnitude of the Balmer jump. These problems are not new. The difficulties in fitting integrated spectra of nova-likes and dwarf nova in outburst with disc model spectra have a long history (e.g., Wade 1984, 1988; La Dous 1989; Long et~al. 1991, 1994; Knigge et~al. 1997). In discussing possible explanations for these problems, Knigge et~al. (1998a) postulated the presence of a significant amount of optically thin material in the system in order to reconcile the disc models with the observed spectrum. Our spatially resolved study confirms their suggestion by revealing that the integrated spectrum of UX~UMa has indeed a substantial contribution from optically thin emission, most probably associated to the uneclipsed parts of the disc chromosphere + wind. A calculation by Knigge et~al. (1998a) indicated that the addition of an optically thin component with $T= 3 \\times 10^4$\\,K, $n_H= 5 \\times 10^{12}\\; {\\rm cm}^{-3}$, and vertical extension $H= 9.7 \\times 10^9$ cm would be enough to bring the combined disc model plus optically thin emission into good agreement with the observed PRISM spectrum. The predicted fluxes of their optically thin component raises from $\\simeq 2$ mJy at 2000 \\AA\\ to $\\simeq 5.2$ mJy at 3600 \\AA, being at the level of $\\simeq 3$ mJy at 4500 \\AA\\ -- in good accordance with the fluxes of the uneclipsed component in Fig.\\,\\ref{ffig6}. For this case, their inferred mass accretion rate is $5 \\times 10^{17}\\; g\\,s^{-1}$ or $10^{-8.1}\\; M_\\odot\\: yr^{-1}$, in excellent agreement with our result. Thus, the reason for the discrepancies between the prediction of the standard disc model and observations is not an inadequate treatment of radiative transfer in the disc atmosphere (or standard models of vertical structure), but rather the presence of additional important sources of light in the system besides the accretion disc (e.g., optically thin continuum emission from the disc wind and possible absorption by circumstellar cool gas). Following the same line of reasoning, if disc winds are a common characteristic of all non-magnetic, high state cataclysmic variables, one might expect their disc chromospheres to contribute a non-negligible amount of optically thin emission to the total light of the system. Under this hypothesis, the discrepancy between disc models and the integrated spectrum observed in other non-magnetic, high state systems may be removed by the inclusion of a proper optically thin component to their total light. These results underscore the importance of spatially resolved studies in disentangling the different components of the integrated spectra of cataclysmic variables. In this particular case, it helped to clarify the situation regarding the apparent discrepancy between disc atmospheres models and the observed spectra." }, "9804/astro-ph9804287_arXiv.txt": { "abstract": "Randich and Schmitt [1995, A\\&A 298, 115] found that the coronal activity of solar-type and low mass stars in Praesepe is significantly lower than that of stars in the Hyades cluster. This result is quite surprising since the Hyades and Praesepe have approximately the same age and metallicity and are often thought to have originated in the same Giant Molecular Cloud complex. We have carried out several tests in order to find a possible explanation for this result. We have measured radial velocities of two groups of Praesepe stars (a dF-dK sample and a dM sample) and have measured H$\\alpha$ as a chromospheric activity index for the dM sample. Based on analyses of these data, we conclude that the Praesepe catalog used in the X-ray analysis does not contain a significant number of non-members, and thus that membership problems do not seem to be the cause of the Randich and Schmitt result. The comparison of the H$\\alpha$ equivalent widths for the M dwarfs in Praesepe with those in the Hyades indicates that, at least for stars in this mass range, the Praesepe stars are as active or more active than their Hyades counterparts. The similarity of chromospheric emission allows us to reject differences in the rotational velocity distribution as the origin of the dissimilar Lx luminosity functions. We have also analyzed a few ROSAT PSPC pointings of Praesepe in order to obtain a new and independent estimate of the X-ray luminosities and upper limits for a small sample of Praesepe members. This analysis suggests that the previous ROSAT/PSPC analysis produced slightly optimistic X-ray upper limits; however, the differences between the old and new upper limits are not large enough to explain the dichotomy in the X-ray properties of Praesepe and the Hyades. Therefore, our examination of the available data does not provide a clear reason to explain why the X-ray luminosity functions of the two clusters are different. Part of the explanation could be found in the binaries. Speculatively, these clusters could have different orbital period distributions, with more short period binaries among the Hyades, which would show larger coronal activity. ", "introduction": "Open clusters play a key role in the understanding of different time-dependent stellar properties such as the evolution of rotation, stellar activity, and the lithium abundance. The comparison between different clusters which have the same age allows us to prove if this approach is correct or if other effects, such as e.g. a different metal content, are also important. In this report, we examine the Hyades and Praesepe clusters. During the last 15 years, it has been demonstrated that X-ray emission is a `normal' characteristic of late type stars. As with other stellar properties which depend on rotation (in this case through the dynamo effect, Parker 1955), the emission strength decays with age, as shown by the comparison between open clusters of different ages such as the Pleiades (Caillault and Helfand 1985; Micela et al. 1985; Micela et al. 1990; Stauffer et al. 1994), and the Hyades (Stern et al. 1981; Pye et al. 1994; Stern et al. 1994; Stern et al. 1995). The Hyades cluster has been extensively studied at X-ray wavelengths. The ROSAT All-Sky Survey (RASS) detected members down to Log~Lx=1--2$\\times$10$^{28}$ erg~s$^{-1}$, with detection rates of 90\\% for spectral type dG, 40\\% for dK and 30\\% for dM stars (Stern et al. 1995). They also found that X-ray luminosity functions (XLDF) of K and M-type dwarfs are significantly affected by the presence of a large number of binary systems in the cluster, in the sense that an important fraction of the stars with strong X-ray emission were binaries. A study of the X-ray properties of Praesepe has been carried out by Randich and Schmitt (1995). The Hyades and Praesepe have similar age and metallicity although the Hyades are slightly more metal rich. Moreover, their kinetic properties are quite close (Eggen 1992) and they could have been born in the same molecular cloud. Randich and Schmitt (1995) presented the results from ROSAT PSPC Raster Scan images in a 4$^\\circ$$\\times$4$^\\circ$ region. Their detection rates for Praesepe were 33\\%, 14\\% and 13\\% for dG, dK and dM stars, respectively. As a consequence, the X-ray luminosity functions of Praesepe in each spectral range are dominated by the upper limits (UL). Since a large fraction of the Praesepe Raster Scan was characterized by a sensitivity similar to that of the Hyades RASS observation, the difference in the detection rates means that the bulk of the Praesepe population is underluminous in X-rays with respect to the Hyades. The goal of this papers is to try to disentangle this problem, looking for possible reasons of the disparate behavior of these coeval clusters in X-rays. We present the Praesepe data studied here and the reduction process in Section 2, where in Section 3 we analyze the data and perform a comparison with the Hyades cluster. Section 4 contains the more important conclusions derived from this study. ", "conclusions": "We have tried to establish the reasons of the different X-ray properties of late type stars in the Hyades and Praesepe. We have studied two different samples of stars: dF-dK Praesepe stars having detected and upper limits for their X-ray luminosities and dM Praesepe stars which have been not detected by ROSAT. The measured radial velocities for both samples show that contamination by spurious members cannot account for the differences in the level of coronal activity, since all stars (but one) studied here, and presumably most of the stars in the Randich and Schmitt (1995) sample, are real members. Using simultaneously color-magnitude diagrams and the measured radial velocities, we have discovered new binaries in Praesepe for the dF-dK stars. The comparison of the fraction of binaries in Praesepe and the Hyades shows that it could be slightly different in both clusters. Since the observed levels of coronal activity, assumed equal sensitivity in the observations, are lower in Praesepe, one would expect a smaller binarity rate in Praesepe than in the Hyades. Moreover, we have shown that the detection rate for the binaries is much higher in the Hyades than in Praesepe. This could be interpreted as an effect of a difference in the distribution of the orbital periods in both clusters. Finally, the study of the statistical properties of the H$\\alpha$ spectral line for the dM stars in both clusters shows that in fact Praesepe presents higher chromospheric activity for this kind of stars than the Hyades. This result is also surprising, since none of the Praesepe M dwarfs were detected in X-ray, whereas many of the Hyades M dwarfs are coronally active. For this reason, one possible explanation for the differences in the X-ray properties between both coeval clusters, the existence of different distributions of the rotational velocities, seems unlikely. We have found several Praesepe dM stars which have a remarkable strong H$\\alpha$ emission and very low Lx upper limits, an unexpected situation. All these data could indicate the possibility of a difference between the sensitivity of the ROSAT All-Sky Survey for the Hyades and the ROSAT PSPC observations of Praesepe. However, our re--analysis of several ROSAT pointings shows that the previous assignation of upper limits is essentially correct (although there is a suggestion that the initial estimates for the upper limits were too low by a factor of two). We propose differences in the orbital period distribution as a partial explanation of the dichotomy of the Lx properties. Extensive studies of different properties which characterize late-type stars, such as rotational velocities and periods, lithium abundances and additional activity indicators should be made in a large variety of open clusters in order to have a comprehensive perspective of the evolution of this type of stars. In particular, a similar comparison to that performed here with other clusters of the same age, such as Coma, could also contribute towards an understanding of the differences in the X-ray properties of coeval clusters. New X--ray data from AXAF or XMM could help to solve this problem." }, "9804/astro-ph9804078_arXiv.txt": { "abstract": "The synchrotron reflection scenario recently proposed to explain $\\gamma$-ray flares observed from blazar jets is studied. Our analysis takes into account the angular distribution of the beamed radiation, the finite extent of the scattering region, and light travel-time effects. We compare energy densities and powers for synchrotron, SSC, reflected synchrotron (RSy), and external Compton (EC) scattering processes. If the width of the scattering layer is much larger than $\\Gamma R^\\prime_B$, where $\\Gamma$ and $ R^\\prime_B$ denote the bulk Lorentz factor and comoving-frame radius of the plasma blob, respectively, then the ratio of the RSy and synchrotron energy densities $\\sim 4 \\, \\Gamma^3 n_{\\rm BLR} \\sigma_{\\rm T} R^\\prime_B$, where $n_{\\rm BLR}$ is the mean particle density in the broad line region (BLR). Our results imply that Thomson-thick scattering regions of narrow extent must be present for the synchrotron reflection mechanism to operate effectively. This process seems unlikely to cause flares in lineless BL Lac sources, where X-ray and TeV flares are common and the BLR is thought to be weak or absent. We sketch time profiles of flares for various scenarios, including a model where the blob is energized by sweeping up surrounding material. ", "introduction": "More than 50 blazar-type AGNs have been detected with high confidence by EGRET to emit $\\gamma$-rays above 100~MeV (Matox et al. \\markcite{Mattox97}1997). These sources are identified with flat-spectrum radio sources classified as BL-Lac objects or quasars. Many of these objects exhibit variability on all wavelengths, with some of the most rapid variability, on time scales of hours to days, observed at the highest $\\gamma$-ray energies (e.g., Bloom et al. \\markcite{bloom97}1997; Wagner et al. \\markcite{wagner95}1995; Mukherjee et al. \\markcite{mukerjee97}1997). The large apparent luminosities in combination with the short variability time scales provide evidence for the widely accepted relativistic jet model for AGNs (for recent reviews, see Schlickeiser \\markcite{schlickeiser96}1996 and Hartman et al. \\markcite{Hartman97}1997), according to which the radio--$\\gamma$-ray emission from blazars is emitted via nonthermal synchrotron radiation and Comptonization of soft photons by energetic particles in relativistic outflows powered by accreting supermassive black holes. Soft photons which are Compton-scattered to produce the $\\gamma$-ray emission include internal synchrotron photons (e.g., Marscher \\& Gear \\markcite{mg85}1985, Maraschi et al. \\markcite{maraschi92}1992, Bloom \\& Marscher \\markcite{bm96}1996) and accretion-disk radiation which enters the jet directly (Dermer \\& Schlickeiser \\markcite{ds93}1993) and after being scattered by surrounding BLR clouds and circumnuclear debris (e.g., Sikora, Begelman \\& Rees \\markcite{sbr94}1994; Blandford \\& Levinson \\markcite{bl95}1995; Dermer, Sturner, \\& Schlickeiser \\markcite{dss97}1997; Protheroe \\& Biermann \\markcite{pb97}1997). It has recently been proposed (Ghisellini \\& Madau \\markcite{gm96}1996; hereafter GM96) that the beamed synchrotron radiation, after scattering off a cloud near the jet trajectory and reentering the jet, can be a source of copious soft photons and lead to a pronounced flare of very short duration as the relativistic jet plasma passes through the cloud. Wehrle et al. (\\markcite{Wehrle98}1998) argue that this mechanism might explain the February 1996 $\\gamma$-ray flare observed from 3C 279. A detailed analysis which correctly accounts for causality effects and the finite width of the scattering layer was not performed by GM\\markcite{gm96}96, and such a treatement is required before blazar flare spectra and light curves can be modeled. Here we examine this model in more detail for a simple geometry of the BLR in the limiting regime where the nonthermal jet electron distribution does not evolve. In \\S 2 we describe the model. Numerical calculations of the magnetic-field and photon energy densities and synchrotron and Compton powers are presented in \\S 3. Application to blazar flares is made and the effects of blob energization on time profiles of flares are indicated. We summarize in \\S 4. ", "conclusions": "Photon energy densities and radiative powers due to different processes were numerically calculated as a function of the distance of the blob from the accretion disk for a wide range of parameters. Figure 2 shows an example of our series of simulations, using parameters representative of a flat-spectrum radio quasar (FSRQ) and a BL Lac object (BL). Here we let $p = 3$ (FSRQ; solid curves) and $p = 2.7$ (BL; shaded curves). This choice yields fairly flat $\\nu F_{\\nu}$ spectra which can be compared with the peaks of the broadband $\\nu F_{\\nu}$ spectral energy distributions of blazars. The electron density and blob radius are chosen so that the resulting total luminosities are in accord with typical values of the apparent luminosities of FSRQs and BLs, and so that $< 100$ GeV $\\gamma$ rays are not absorbed by $\\gamma$-$\\gamma$ pair production on the synchrotron photons intrinsic to the source. For the generic FSRQ, we assume that the BLR has a radial Thomson depth $\\tau_{\\rm T, BLR} = 0.2$ and occupies a spherical shell located between 0.05 and 0.5~pc from the central engine. For simplicity, we assume that the accretion disk radiates isotropically with a luminosity of $10^{46}$~erg~s$^{-1}$. This choice of parameters gives a Compton power which is $\\sim 10$ times greater than the synchrotron power. Figure 2 shows that although the RSy radiation energy density $u'_{RSy}$ is larger than $u'_{Sy}$ when the blob is located within the BLR, the RSy radiative power is about equal to the SSC power. This is because the backscattered jet synchrotron photons are boosted in energy by a factor of $\\Gamma^2$ relative to the energy of the synchrotron photons in the comoving frame; thus Klein-Nishina effects reduce the radiative power resulting from Compton scattering of RSy radiation more strongly than for the SSC radiation. In our simulations using plausible parameters compatible with the assumed spherical shell geometry of the BLR, we find that the Compton power due to the synchrotron mirror mechanism is at most comparable to the SSC power. The efficiency of this process is improved when $\\Gamma \\gg 10$ and a very narrow ($\\Delta r_{\\rm BLR} \\lesssim r_{\\rm in} / (2\\Gamma^2)$), Thomson-thick BLR is located very far ($r_{\\rm in} \\gg \\Gamma^2 R'_B$) from the central engine. The synchrotron reflection process might therefore operate in BLR clouds which are thought to surround Seyfert AGNs and FSRQs. BLR clouds, as understood through photoionization models (see, e.g., Wandel \\markcite{Wandel97}1997 and references therein), consist of dense ($n\\sim 10^{10}-10^{11}$ cm), Thomson-thick ($\\tau \\sim 1-10$) regions covering a small ($\\sim 10$\\%) fraction of the central engine. For such a model to be feasible, however, the conditions regarding duty cycle and power outlined by Dermer \\& Chiang (\\markcite{dc98}1998) must be met. We defer presentation of results in the regime $R^\\prime_B \\gtrsim \\Delta r_{\\rm BLR}/\\Gamma$ to future work; the shell geometry used here is excessively artificial in this limit. The absence of strong emission lines in X-ray selected BLs and (to a lesser extent) radio-selected BLs suggests that the BLR is considerably more dilute in BLs than in FSRQs, and that BLs have mean Thomson thicknesses $\\tau_{\\rm T,BLR} \\ll 0.1$ (see Scarpa \\& Falomo \\markcite{sf97}1997 and references therein). (On the other hand, the strength of the central ionizing photon source might be much less in BLs than FSRQs.) Superluminal motion observations also indicate that typical values of $\\Gamma$ for BLs lie in the range between $\\sim 3$ and 7 (see the review by Urry \\& Padovani \\markcite{up95}1995). The ability of the synchrotron reflection process to produce gamma-ray flares therefore seems more difficult in BLs than in FSRQs, yet TeV flares often coincident with X-ray flares have been detected from three BL Lac objects (e.g., Punch et al. \\markcite{punch92}1992, Macomb et al. \\markcite{Macomb95}1995; Catanese et al. \\markcite{Catanese97}1997, \\markcite{Catanese98}1998). For the assumed BL parameters in Figure 2, the synchrotron reflection flare could hardly be detected. If an accretion disk steadily radiates photons (see B\\\"ottcher \\& Dermer \\markcite{bd95}1995 for a treatment of time-variable disk radiation), then light-travel time effects can be neglected. The accretion disk supplies an abundant supply of soft photons, which can enter the jet directly (ECD) and after being scattered by the BLR (ECC). The comoving photon energy densities from the ECD process can dominate that from the ECC and synchrotron processes when $z \\lesssim 10^{-2}$~pc, but declines $\\propto z^{-3}$ and $\\propto z^{-2}$ farther out (see Dermer \\& Schlickeiser \\markcite{ds93}1993; B\\\"ottcher, Mause \\& Schlickeiser \\markcite{bms97}1997). The ECC photon energy density increases slowly with $z$ when $z < r_{\\rm in}$, and begins to decrease for $r_{\\rm in} \\lesssim z \\lesssim r_{\\rm out}$. Outside the BLR, when $z\\gtrsim r_{\\rm out}$, the energy density asymptotically approaches the limiting behavior $u^\\prime_{\\rm ECC} \\propto z^{-2}$. When $z \\lesssim r_{\\rm in}$, the ECC process dominates over the SSC process provided that \\begin{equation} {\\tau_{\\rm T, BLR} \\over r_{\\rm in}^2} \\gtrsim {2 \\over 3} {c \\, B^2 \\tau_B \\over L_D \\, \\Gamma^2} \\> \\big( {p - 1 \\over 3 - p}\\big) \\> {\\gamma_2^{3 - p} - \\gamma_1^{3 - p} \\over \\gamma_1^{1 - p} - \\gamma_2^{1 - p}}, \\end{equation} where $\\tau_B = R^\\prime_B \\, n_{\\rm e,jet} \\, \\sigma_T$ is the Thomson depth of the blob. Using the parameters adopted in Figure 2 but letting $\\tau_{\\rm T, BLR}$, $r_{\\rm in}$, and $L$ vary, we find that the ECC photon energy density dominates the synchrotron photon energy density in the comoving blob frame when $ r_{\\rm in}({\\rm pc}) \\lesssim 0.4 \\> \\tau^{1/2}_{\\rm T, BLR} \\, L_{46}^{1/2}$ and $ r_{\\rm in}({\\rm pc}) \\lesssim 0.1 \\> \\tau^{1/2}_{\\rm T, BLR} \\, L_{44}^{1/2}$ for the FSRQ and BL parameters, respectively, where $L_n = L_{\\rm disk} / (10^n~{\\rm erg \\> s}^{-1})$. It should be noted that Klein-Nishina effects are not included in estimate (6). The bottom panel of Figure 2 illustrates the time profiles of a flare calculated using the FSRQ parameters for the ECC (thick solid curve) and RSy (thick dot-dashed curve) processes. Note that the observer's time element is linearly related to $z$ for a blob moving with constant velocity. The ECC process gives a fast-rise, power-law-decay--type light curve, and the RSy mechanism gives a gradual rise of the $\\gamma$-ray flux and a sharp drop as the blob leaves the BLR. A flare produced by the RSy process could be identified by a rapid decline of $\\gamma$-rays which is not accompanied by a corresponding decrease of the synchrotron emission. For comparison, we also sketch a flare time profile produced by the ECD process, and time profiles produced by sweeping energization of the blob. In this process, the bulk kinetic energy of the outflowing plasmoid is converted into internal nonthermal particle energy by sweeping up BLR material (see, e.g., Panaitescu \\& M\\'esz\\'aros \\markcite{pm98}1998; Dermer \\& Chiang \\markcite{dc98}1998; Chiang \\& Dermer \\markcite{Chiang98}1998). The pair of light solid and dot-dashed curves illustrate ECC and RSy flares, respectively, are modeled assuming that the nonthermal electron energy is proportional to the amount of swept-up matter which is then added to a nonthermal lepton distribution accelerated at the base of the jet. Blob deceleration is assumed to be negligible here. The upper curves of the two pairs are modeled assuming no radiative losses, and the lower curves of the two pairs illustrate the effects of radiative losses on the nonthermal leptons by crudely multiplying the upper curves by a decaying exponential with a $1.5\\times 10^{18}$~cm decay length. The model RSy light curves, with and without sweeping energization, are similar to $\\gamma$-ray light curves observed in the 1991 and 1996 flares of 3C 279 (Hartman et al. \\markcite{Hartman96}1996; Wehrle et al. \\markcite{Wehrle98}1998). This may be a consequence, however, of the highly idealized BLR geometry used in the calculation. More symmetical flaring profiles observed from PKS 0528+134 (Collmar et al. \\markcite{collmar97}1997), PKS 1622-297 (Mattox et al. \\markcite{Mattox97}1997), and PKS 1406-076 (Wagner et al. \\markcite{Wagner95}1995) might be more easily explained by the ECD or ECC processes. The declines of the X-ray and optical fluxes correlated with the EGRET $\\gamma$-ray fluxes in the February 1996 3C 279 and the 1406-076 flares could, however, rule out the RSy mechanism since the synchrotron component is not directly affected by the reflection process. The inclusion of electron energy evolution and the relaxation of the assumption of a constant velocity blob must be treated to strengthen such conclusions. Future flare modeling must treat the passage of the jet through the BLR clouds, which themselves are in Keplerian motion around the central black hole. The passage of a jet through such a region will display a complicated signature when monitored by different telescopes with different sensitivities and imaging capabilities, which can only be decoded when full account is taken of the processes considered here. The efficiency of the different Compton-scattering scenarios, including the RSy mechanism, in this more realistic system is presently under investigation by the authors." }, "9804/astro-ph9804308_arXiv.txt": { "abstract": "We consider the implications of the detection of spiral structure in the accretion disc of the binary IP Pegasi. We use numerical simulations of the development of a disc outburst to construct predicted Doppler tomograms, which are found to be in close agreement with the observations if the spiral pattern arises as a transient feature when the disc expands viscously at the start of the outburst. The good agreement of such viscous disc simulations with the data is consistent with models in which most of the angular momentum transport in the disc originates in internal stresses rather than globally excited waves or shocks. Future detailed observations of the development of transient spiral features offer the potential to measure the dependence of the disc viscosity on the local physical conditions in the disc. ", "introduction": "Recent observations of the dwarf nova IP Pegasi provide convincing evidence for spiral structure in the emission from an accretion disc in a binary system (Steeghs, Harlaftis \\& Horne 1997). During an outburst, changes in the profile of spectral lines with binary phase were inverted using the technique of Doppler tomography (Marsh \\& Horne 1988) to reveal a loosely wrapped, two-armed spiral pattern in the disc emission. No such structure is observed in the quiescent disc (Marsh \\& Horne 1990). These observations provide a potential new constraint on the angular momentum transport processes operating in accretion discs. Two mechanisms are known that can provide a source of viscosity in ionized, non-self-gravitating accretion discs in binary systems; turbulence driven by the non-linear development of the Balbus-Hawley instability (Balbus \\& Hawley 1991; Tout \\& Pringle 1992; Stone et al. 1996; Brandenburg et al. 1996); and spiral waves or shocks driven by the gravitational perturbation of the secondary (Sawada, Matsuda \\& Hachisu 1986; Spruit 1987; Rozyczka \\& Spruit 1989; Savonije, Papaloizou \\& Lin 1994). It is obvious that the second scenario leads to a spiral pattern of disc emission, but even if tidally induced shocks are unimportant for the angular momentum transport budget in the steady-state they might still be observable in outburst, when the enhanced viscosity forces the disc to expand into a region where the strength of the tidal forces is greater (Papaloizou \\& Pringle 1977; Lin \\& Pringle 1976). We note that although it is generally believed that the spiral shock mechanism is inefficient in the relatively cool discs found in cataclysmic variables (Livio 1994; Savonije, Papaloizou \\& Lin 1994), there are considerable theoretical uncertainties in both mechanisms, and additional observational input is highly desirable. In this Letter, we compare simulations of accretion disc evolution with the observations of IP Peg. Our goal is to test whether viscous disc simulations, which are predicated on the existence of an internal origin for the disc viscosity, are consistent with the strong spiral structure observed in the data. We describe our calculations in Section 2, and present in Section 3 model Doppler tomograms for comparison with the observations. Section 4 summarises our conclusions, and outlines the theoretical expectations for spiral structure in other disc systems. ", "conclusions": "In this Letter we have presented a simulation of the evolution of the accretion disc for the parameters of the binary IP Pegasi in outburst. We find as our main result that a spiral pattern is formed in the outer disc during outburst as the enhanced viscous stresses push the disc edge into a region of strong gravitational perturbations from the secondary. The spiral structure obtained in the simulation is two-armed, non-resonant, and much more prominent in outburst as compared to quiescence. Comparing the results with the observations of Steeghs, Harlaftis \\& Horne (1997), we find that there is excellent agreement with the azimuthal extent, velocity range, and asymmetry of the observed pattern. The observations of IP Pegasi and other cataclysmic variables find no clear evidence for spiral patterns in quiescent discs. This is consistent with theoretical expectations provided that the quiescent discs are cool, viscous, and have not expanded close to the tidal radius. Observations following the decay of spiral structure after the outburst has ended would be valuable in understanding the interplay of these factors. However the current evidence continues to support the conclusion of Savonije, Papaloizou \\& Lin (1994) that angular momentum transport by spiral shocks is insufficiently effective to provide the bulk of the angular momentum transport in the relatively cool, thin discs found in cataclysmic variables. Spiral shocks {\\em are} likely to be important over a wide range of radii in the hotter discs found in X-ray binaries (Owen \\& Blondin 1997), and in the quasi-spherical accretion flows postulated for the advectively dominated regime, although internal sources of viscosity are also likely to be more efficient in those thicker disc geometries. The current observations can be modelled adequately using a highly simplified three-dimensional model of an outburst caused by a thermal disc instability, in which the only inputs are the change in $\\alpha_{\\rm SS}$ and $c_s$ between quiescence and outburst. Future observations, extending over the rise to outburst and during the decline, may be able to provide stronger constraints on the assumed disc model. In particular, since the spiral pattern arises as a result of the imbalance between internal viscous stresses and well-understood gravitational torques, such observations can probe the variation of the disc viscosity with the local physical conditions in the disc. Eclipsing systems such as IP Peg are particularly promising in this regard as the radial run of quantities such as the effective temperature can readily be derived simultaneously from eclipse mapping. The detection of spiral structure in the accretion disc of IP Peg, which has relatively feeble outbursts, implies that similar or stronger features may be expected in most dwarf novae. Theoretically, we note that qualitative differences are expected in systems with low mass ratio (roughly, $q < 1/4$). In these binaries both the observations of superhumps in the light curve, and numerical simulations (Murray 1996, 1998), suggest that a strong $m=1$ mode is excited in outburst. These features are {\\em not} stationary in the corotating binary frame, making detailed investigation via Doppler mapping harder. It would also be worthwhile investigating whether spiral structure is observable in magnetic systems such as intermediate polars, where non-axisymmetry might be induced at the {\\em inner} edge of the disc as a result of magnetic torques from the white dwarf. {\\em Note added:} Godon, Livio \\& Lubow (1998) have recently presented calculations showing that a steady tidally induced spiral pattern does not match the observations of IP Peg. This is consistent with our finding that consideration of the transient behaviour of the disc during outburst is required." }, "9804/astro-ph9804293_arXiv.txt": { "abstract": "The Carina Nebula is an extremely bright southern \\HII\\ region embedded in a giant molecular cloud and contains some of the most massive stars known in our Galaxy. We are undertaking a multi-wavelength study of the Carina Nebula in order to examine the detailed kinematics and distribution of the molecular and ionised gas, and to look for further evidence of ongoing star formation. Here we present the results of the initial molecular cloud observations which were made by observing the \\CO\\ emission with the Mopra antenna. The observations reveal the clumpy morphology of the molecular gas, and allow us to identify many interesting regions for follow-up observations. ", "introduction": "The Carina \\HII\\ region/molecular cloud complex is an excellent region for studying the interaction of massive stars with their parental Giant Molecular Cloud (GMC). The nebula covers an area of $\\approx$ 4 deg$^{2}$ and is bisected by a prominent V-shaped dark lane. There are over 14 star clusters in this region which have been studied extensively over the past twenty years. For excellent reviews see Feinstein (1995) and Walborn (1995). The most influential clusters of the nebula are the two OB clusters, Tr 14 and Tr 16. These clusters contain numerous O-type stars, including three O3 stars each, making them two of the most massive star clusters in our galaxy. Tr 14 is a compact cluster situated to the north-west of the nebula, adjacent to the western dust lane. Tr 16 is an open cluster centred northwards of the vertex of the dark lane. It contains one of the most massive stars known: $\\eta$ Car. Here we will adopt the popular view (e.g. Tovmassian 1995, Walborn 1995) that Tr 14 and Tr 16 are at a common distance of about 2.2 kpc and that Tr 14 is younger than Tr 16. Considering the extensive studies on the stellar content of the Carina Nebula, in particular $\\eta$ Car and its surrounding Homunculus nebula, relatively little work has been done on the extended nebula in the last fifteen years. Early radio continuum observations revealed that the nebula contains a large ionised region with two peaks, Car I and Car II (Gardner \\& Morimoto 1968). Higher resolution radio continuum data show that both Car I and Car II are made up of a number of filamentary arcs and rings and are everywhere thermal (Retallack 1983, Whiteoak 1994). Car II is located to the north of $\\eta$ Car and Car I is located towards the western dark lane, just west of Tr 14. The dynamics of the ionised gas in this region have been studied via hydrogen recombination line emission (Gardner et al. 1970, Huchtmeier \\& Day 1975) and H$\\alpha$ and [\\NII] emission observations (Deharveng \\& Maucherat 1975). The results show line splitting towards the Car II region which has been interpreted as an expanding shell of ionised gas. The dark lanes consist of molecular gas and dust that are associated with the nebula (Dickel 1974). H$_{2}$CO and OH absorption measurements identified two optical depth maxima which were located towards these lanes (Gardner, Dickel \\& Whiteoak 1970, Dickel \\& Wall 1974). Extended far-IR emission is confined there also (Harvey Hoffmann \\& Campbell 1979, Ghosh et al. 1988). There are two main CO emission regions towards the nebula; a northern and southern cloud (de Graauw et al. 1981, Whiteoak \\& Otrupcek 1984). Both regions are part of a much larger GMC which has a projected length of 130 pc and a mass in excess of 5$\\times$10$^{5}$ \\Msun (Grabelsky et al. 1988). The area between the southern and northern CO clouds is centred on the Keyhole Nebula, a dense dark cloud northwest of $\\eta$ Car. Here the molecular gas exists in dense clumps of typical mass 10 \\Msun\\ that are separated both in space and velocity (Cox \\& Bronfman 1995). The picture used to describe the Carina complex is one in which the massive star clusters, Tr 14 and Tr 16, are interacting strongly with the molecular cloud from which they formed. It is generally accepted that the photons from Tr 14 and Tr 16 are responsible for the ionised emission of Car I and Car II respectively, and that their strong stellar winds are producing the general expansion of the nebula. We are undertaking a multi-wavelength study of the Carina Complex in order to study the detailed kinematics and distribution of the molecular and ionised gas and to look for further evidence of ongoing star formation. Here we present the results of initial observations of the \\CO\\ emission. CO emission is thermalised in both low- and high-density gas and therefore is suitable for tracing the overall distribution and velocity structure of the molecular cloud. It also can pinpoint any `CO hot-spots'. These are warm molecular cores where stars could possibly be forming. ", "conclusions": "We have presented data from the first stage of an extensive study of the GMC associated with the Carina nebula. The data consists of observations of \\CO\\ emission which have been used to trace the overall GMC as well as pinpoint and CO `hot-spots' or dense regions where stars could possible form. The observations are at a higher resolution than previous studies and reveal the clumpy nature of the northern and southern cloud regions. They also show the positional coincidence between the far-infrared emission and the strong CO emission towards the Car I region. This supports a blister-type model for this region. Further observation of different transitions are being made to better constrain the temperature and density of the molecular gas in this interesting region." }, "9804/gr-qc9804051_arXiv.txt": { "abstract": "Linde's proposal of a Euclidean path integral with the ``wrong'' sign of Euclidean action is often identified with the tunneling proposal for the wave function of the universe. However, the two proposals are in fact quite different. I illustrate the difference and point out that recent criticism by Hawking and Turok does not apply to the tunneling proposal. ", "introduction": " ", "conclusions": "" }, "9804/astro-ph9804016_arXiv.txt": { "abstract": "We report on a near-infrared, long-baseline interferometric search for luminous companions to the star 51~Pegasi conducted with the Palomar Testbed Interferometer. Our data is completely consistent with a single-star hypothesis. We find no evidence to suggest a luminous companion to 51~Pegasi, and can exclude a companion brighter than a $\\Delta$K of 4.27 at the 99\\% confidence level for the 4.2-day orbital period indicated by spectroscopic measurements. This $\\Delta$K corresponds to an upper limit in the companion M$_K$ of 7.30, in turn implying a main-sequence companion mass less than 0.22 M$_{\\sun}$. ", "introduction": "The recent inference of a planetary-mass gravitational companion to the star 51~Pegasi (HD 217014) from apparent radial velocity variation by Mayor \\& Queloz (1995) has subjected this otherwise unremarkable star to remarkable scrutiny. The Mayor and Queloz result was quickly verified by several groups with similar or higher-resolution spectroscopic techniques (c.f.~\\cite{Marcy97}). However, there has been no other evidence for a companion, e.g.~precision photometric monitoring has failed to show evidence for eclipses (\\cite{Henry97}), and there is a significant lack of x-ray flux from the system compared to binary systems with similar periods (\\cite{Pravdo96}). Further, 51~Peg's G5V spectral classification has become mildly controversial (e.g.~\\cite{Houk95}, who argues for a G2-3V), as has its physical size (e.g.~\\cite{Hatzes97,Henry97}). A planetary-mass companion in a 4.2 day orbit around a solar-mass 51~Peg would have an orbital semi-major axis of approximately 0.05 AU (\\cite{Marcy97}), slightly more if the companion were more massive. At a distance of 15.4 $\\pm$ 0.2 pc (\\cite{Perryman96}), the approximate maximum primary-companion angular separation would be 3.5 millarcseconds (mas). Such an angular separation is well below resolution limits for current conventional imaging technology, but is accessible to optical and near-infrared interferometry. As only the lower mass limit is set by the spectroscopic results, it is possible the companion is significantly more massive -- perhaps even a low-mass star. We have therefore studied 51~Peg with the Palomar Testbed Interferometer (PTI) in an attempt to detect the putative companion if it is indeed sufficiently luminous. PTI is a 110m-baseline interferometer operating at K-band (2 -- 2.4 $\\mu$m) located at Palomar Observatory, and described in detail elsewhere (\\cite{Colavita94}). The minimum PTI fringe spacing is roughly 4 mas at the sky position of 51~Peg, making a (sufficiently) luminous companion readily detectable. ", "conclusions": "We find no evidence to suggest that the putative 4.2-day period companion to 51~Peg is detectable in our data; all of the datasets we have analyzed indicate that 51~Peg is at least as stable as our two calibration sources. The 1997 PTI data on 51~Peg is sufficiently stable that we can place significant limits on $\\Delta$K and consequently M$_K$ of a 4.2-day period companion. We find upper limits in $\\Delta$K of 4.78, 4.53, and 4.27 for the 4.2-day period companion to 51~Peg at 68\\%, 95\\%, and 99\\% confidence levels respectively. These $\\Delta$K limits imply companion M$_K$ limits of 7.81, 7.56, and 7.30, corresponding to upper limits on the mass of a putative main sequence companion at 0.17, 0.20, and 0.22 M$_{\\sun}$ at the 68\\%, 95\\%, and 99\\% confidence levels respectively (\\cite{Henry93}). Our results cannot exclude the possibility of a very low-mass star in a face-on orbit as the 51~Peg companion, but such a star would have to be of spectral type M5V or later." }, "9804/astro-ph9804220_arXiv.txt": { "abstract": "We numerically investigate stellar and gas dynamics in star-forming and dissipative galaxy mergers between two disk galaxies with specific orbital configurations. We find that violent relaxation combined with gaseous dissipation in galaxy merging transforms two disk galaxies into one S0 galaxy with polar-rings: Both the central S0-like host and the polar-ring component in a polar-ring galaxy are originally disk galaxies. We also find that morphology of the developed polar-rings reflects both the initial orbit configuration of galaxy merging and the initial mass ratio of the two merger progenitor disk galaxies. Based upon these results, we discuss the origin of the fundamental observational properties of polar-ring galaxies, such as the prevalence of S0 galaxies among polar-ring galaxies, the rarity of polar-ring galaxies among S0 galaxies, the dichotomy between narrow polar-rings and annular ones, shapes of polar-ring warps, and an appreciably larger amount of interstellar gas in the polar-ring component. ", "introduction": "Polar-ring galaxies are generally considered to be dynamically peculiar systems in which the outer rings composed of gas and stars are aligned roughly in a perpendicular orientation with respect to the major axis of the central host galaxies (Schweizer, Whitmore, \\& Rubin 1983; Whitmore et al. 1990; Sackett 1991). A growing number of observational studies have been recently accumulated which can provide valuable information about the origin of these peculiar polar-ring galaxies. Nearly all of the central host are morphologically normal S0 galaxies, some of which are confirmed to be rapidly rotating by kinematical studies (Schechter \\& Gunn 1978; Schechter, Ulrich, \\& Boksenberg 1984; Whitmore et al. 1990; Whitmore 1991). Approximately only 0.5 percent of all S0 galaxies have observable polar-rings, which suggests that a particular mechanism is required for the formation of polar-rings in S0 galaxies. The ring component also shows rapid rotation comparable to that of the main host galaxies, implying that two dynamically different system coexist in these polar-ring galaxies. An appreciable amount of HI gas, which is sometimes comparable to the total mass of the host, is closely associated with the stellar ring component (e.g., Shane 1980; Schechter et al. 1984; Richter, Sackett, \\& Sparke 1994; Arnaboldi et al. 1997; Galletta, Sage, \\& Sparke 1997). The morphology of the polar-rings is basically divided into two broad classes (Whitmore 1991): A narrow ring which is not extended in size (e.g., ESO 415-G 26) and an annulus which is a disk-like component with the central part cut out (e.g., NGC 4650a). Peculiar morphology is observed in some polar-ring galaxies (e.g., the Helix galaxy, NGC 2685, and double ringed system, ESO 474-G 26), which further implies considerably complicated physical processes in polar-ring formation and simultaneously provides a clue to the understanding of the origin of polar-ring galaxies (Sackett 1991). Roughly two-thirds of these polar-rings show obvious galactic warps whose shapes look like `integral sign' and/or `banana'(Whitmore 1991). Statistical studies on the distribution of the angle between the ring component and the central host reveal that these two components strongly prefer to be orthogonal with each other. These peculiarities both in the kinematics and morphology observed in polar-ring galaxies have attracted a number of theoretical interests, which are divided basically into two categories: One is the origin of the polar-rings and the other is the nature of dark matter halo surrounding polar-ring galaxies. Although there are a large number of important studies addressing the three dimensional shapes of dark matter halo in the galaxies (e.g., Whitmore, Mcelroy, \\& Schweizer 1987; Reshetnikov \\& Combes 1994; Sackett et al. 1994; Combes \\& Arnaboldi 1996), we here restrict ourselves to the mechanisms which would naturally explain the formation of the polar-ring galaxies with spherical haloes. It is generally believed that the formation of the polar-rings is the results of a `secondary event' involving a pre-existing S0 galaxy (e.g., Steiman-Cameron \\& Durisen 1982; Sparke 1986; Quinn 1991; Rix \\& Katz 1991; Reshetnikov \\& Sotnikova 1997). Specifically, the host S0 galaxy is supposed to have acquire the material constituting the ring component by capturing the gas during tidal interaction with neighbor galaxies. The subsequent gravitational interaction combined with the gaseous dissipation then spreads the captured gas and forms the polar-rings around the host galaxy. One of the promising models along this orthodox scenario is the `preferred plane model' in which the differential precession of the rings and the gaseous dissipation cooperate to play a vital role in leading the acquired gas to settle into the stable polar orbit and finally to form the polar-rings (Tohline \\& Osterbrock 1982; Durisen et al. 1983; Schweizer et al 1983). A number of numerical simulations have already confirmed in what physical conditions the polar-rings are more likely to form and continue to exist for a relatively longer time-scale (Habe \\& Ikeuchi 1988; Christodoulou et al. 1992; Katz \\& Rix 1992). Indeed these previously proposed models have provided a potential success in reproducing the polar-rings in S0 galaxies, however, these seem to be incapable of giving sufficiently conclusive and persuasive answers to the following seven questions on the origin of the polar-ring galaxies (Sackett 1991; Whitmore 1991; Arnaboldi et al. 1997; Galletta et al. 1997): (1) Why are nearly all the central host galaxies morphologically classified as S0 ? (2) Why are polar-ring galaxies so rare among S0 galaxies ? (3) Why do some polar-ring galaxies have a narrow ring and some have annuli ? (4) Why are the mass and angular momentum of the ring component comparable or sometimes larger than those of the host ? (5) Why are the rings so `polar' ? (6) Why some polar-rings have considerably peculiar morphology such as helical and double-ringed shapes ? (7) Why is there an appreciably greater amount of interstellar gas in the polar-ring component ? In particular, (1), (4), and (7) could not be explained simply by the previous theoretical models, implying either that more elaborated and sophisticated models along the above orthodox scenario should be considered or that the alternative model should be proposed for the explanation of the above questions. The purpose of this paper is to explore the origan of polar-ring galaxies and to propound a new mechanism which more naturally and reasonably explains the aforementioned observational trends of polar-ring galaxies. In the present study, we consider that the dissipative galaxy merging between two disks is a promising mechanism that quite reasonably answers the above seven questions. Therefore we investigate how the dissipative galaxy merging transforms two disks into one early-type S0 galaxy with polar-ring. Furthermore, we investigate how the orbit configuration of galaxy merging and initial mass ratio of the two progenitor disk galaxies can affect the morphology of polar-rings developed after galaxy merging. In this paper, the galaxy merging with specific orbit configurations and sufficient amount of gaseous dissipation is demonstrated to play a vital role in forming both the cental S0-like host and the surrounding polar-ring component in polar-ring galaxies. This paper is an extended version of Bekki (1997) in which the basic mechanism of polar-ring S0 galaxy formation in galaxy mergers is briefly summarized. The layout of this paper is as follows. \\S 2 describes numerical models for dissipative galaxy merging. \\S 3 gives the results obtained in the present study. In \\S 4, we mainly discuss whether or not the model proposed in the present paper can become a new promising model which naturally and reasonably explains the observational properties of polar-ring S0 galaxies. ", "conclusions": "The present numerical study provides a new mechanism by which both the central S0-like host and the ring component in a polar-ring galaxy are simultaneously formed. Although uncertainties of the numerical treatment of gas dynamics and star formation still remain, it appears that our model has succeeded in reproducing $some$ polar-ring galaxies and explaining naturally a number of important observational properties of them. In the proposed model, the formation of polar-ring galaxies is essentially ascribed to the details of dynamics of dissipative galaxy merging with specific orbital configurations. Specifically, the central host of a polar-ring galaxy is the galaxy which has been inevitably transformed from a late-type spiral into an early-type S0 galaxy during the merging. The ring component, on the other hand, is the `galaxy' which has been dramatically transformed from a late-type spiral into a narrow ring or annuli owing to the violent gravitational interaction and gaseous dissipation during the merging. Although both specific orbital configurations and gaseous dissipation in galaxy merging are required for the formation of polar-ring galaxies in the present model, these constraints also give natural explanations to observed trends such as the prevalence of S0 among polar-ring galaxies (e.g., Whitmore 1990), the rarity of the polar-ring galaxies among S0 galaxies (e.g., Whitmore et al. 1991), and an appreciably larger amount of interstellar gas in polar-rings (e.g., Sackett 1991). Moreover it is found that the morphology of polar-rings such as a narrow ring (e.g., ESO 415 - G 26), annular rings (e.g., NGC 4650a), helical rings (e.g., NGC 2658), and double rings (e.g., ESO 474 - G 26) can reflect both the orbital parameters of galaxy merging and the initial mass ratio of two merger precursor galaxies. Thus, the present study demonstrates that a merger remnant of a gas-rich galaxy merger is one of promising candidates of polar-ring S0 galaxies." }, "9804/astro-ph9804150_arXiv.txt": { "abstract": "We present a new, fully covariant and manifestly gauge-invariant expression for the temperature anisotropy in the cosmic microwave background radiation resulting from scalar perturbations. We pay particular attention to gauge issues such as the definition of the temperature perturbation and the placing of the last scattering surface. In the instantaneous recombination approximation, the expression may be integrated up to a Rees-Sciama term for arbitrary matter descriptions in flat, open and closed universes. We discuss the interpretation of our result in the baryon-dominated limit using numerical solutions for conditions on the last scattering surface, and confirm that for adiabatic perturbations the dominant contribution to the anisotropy on intermediate scales (the location of the Doppler peaks) may be understood in terms of the spatial inhomogeneity of the radiation temperature in the baryon rest frame. Finally, we show how this term enters the usual Sachs-Wolfe type calculations (it is rarely seen in such analyses) when subtle gauge effects at the last scattering surface are treated correctly. ", "introduction": "The calculation of the primary temperature anisotropy in the cosmic microwave background radiation (CMB) resulting from density perturbations has a long history, beginning with the seminal paper by Sachs and Wolfe~\\cite{sachs67}. Since the original Sachs-Wolfe estimate, a wealth of detailed predictions for the anisotropies expected in various cosmological models have been worked out. The calculations are straightforward in principle, but, like many topics in cosmological perturbation theory, are plagued by subtle gauge issues~\\cite{stoeger95a}. The problems of gauge-mode solutions to the linear perturbation equations and the gauge-ambiguity of initial conditions can be eliminated by working exclusively with gauge-invariant variables, as in the widely used Bardeen approach~\\cite{bardeen80} and the less well known covariant approach advocated by Ellis and coworkers~\\cite{ellis89a,ellis89b}. However, gauge issues still arise in connection with the definition of the temperature perturbation and the placement of the last scattering surface~\\cite{stoeger95a,ellis-er97}. The latter gauge issues do not arise at first-order in numerical calculations which integrate the Boltzmann equation in a perturbed universe, since the visibility function (which determines the position of the last scattering surface) multiplies first-order variables giving only a second-order error from the use of a zero-order approximation to the visibility~\\cite{LC-scalcmb}. However, this is not always the case in Sachs-Wolfe style analyses, which integrate along null geodesics back to the surface of last scattering, unless care is taken to ensure that the final result involves only first-order variables on the last scattering surface, which then only need be located to zero-order. In this paper, we present a new expression for the CMB temperature anisotropy arising from linear scalar perturbations which is fully covariant and manifestly gauge-invariant. We obtain our expression by integrating the covariant and gauge-invariant Boltzmann equation~\\cite{LC-scalcmb,LC97-er} along observational null geodesics, paying careful attention to the gauge issues discussed above. Unlike some covariant results in the literature (see, for example,~\\cite{LC97-er,dunsby96b}), the expression derived here can be integrated trivially, in the instantaneous recombination approximation, up to a Rees-Sciama term in universes with arbitrary matter descriptions. (The covariant results in~\\cite{LC97-er,dunsby96b} can only be integrated in baryon-dominated universes, thus excluding CDM dominated universes, and other such models favoured by observation.) We base our treatment on the physically appealing covariant and gauge-invariant formulation of perturbation theory, as described in~\\cite{ellis89a,ellis89b}. In this approach, one works exclusively with gauge-invariant variables which are covariantly-defined and hence physically observable in principle. The covariant method has many advantages over other gauge-invariant approaches (such as that formulated by Bardeen~\\cite{bardeen80}). Most notably, the covariant variables have transparent physical definitions which ensures that predictions are always straightforward to interpret physically. Other advantages include the unified treatment of scalar, vector and tensor modes, a systematic linearisation procedure which can be extended to consider higher-order effects (the covariant variables are exactly gauge-invariant, independent of any perturbative expansion), and the ability to linearise about a variety of background models, such as Friedmann-Robertson-Walker (FRW) or Bianchi models. For universes which are baryon-dominated at last scattering, our expression for the temperature anisotropy may be compared to other gauge-invariant analytic results in the literature. We show that, with suitable approximations, the result derived here reduces to that given by Panek~\\cite{panek86} and corrects a similar result given by Dunsby recently~\\cite{dunsby96b}. For the baryon-dominated universe, we use numerical results for the covariant, gauge-invariant variables on the last scattering surface, obtained from a gauge-invariant Boltzmann code~\\cite{LC-scalcmb}, to discuss the different physical contributions to the primary temperature anisotropy. In particular, we show that on intermediate and small scales, the ``monopole'' contribution to the temperature anisotropy is described by the spatial gradient of the photon energy density, in the energy-frame, on the last scattering surface. Since the (real) last scattering surface is approximately a surface of constant radiation temperature (so that recombination does occur there), the inhomogeneity of the radiation energy density in the energy-frame determines a distortion of the last scattering surface relative to the surfaces of simultaneity in the energy-frame. The extra redshift (due to the expansion of the universe) which the photons incur due to the distortion is seen as a ``monopole'' contribution to the temperature anisotropy on intermediate scales. There is a significant ``dipole'' contribution to the anisotropy on intermediate and small scales, which we discuss also. We end with a discussion of the gauge issues inherent in the original Sachs-Wolfe calculation of the CMB anisotropy~\\cite{sachs67}, focusing on the ``monopole'' contribution to the temperature anisotropy on intermediate scales, described above. This contribution is often missed in Sachs-Wolfe type calculations through an incorrect treatment of gauge effects at the last scattering surface~\\cite{ellis-er97}. (Equivalently, the term is often missed through a failure to recognise the direction-dependence of the ``expected temperature'' used to define the temperature perturbation in many calculations.) This often neglected term, which is not important on large scales, is an essential component of the Doppler peaks in the CMB power spectrum. We employ standard general relativity and use a $(+---)$ metric signature. Our conventions for the Riemann and Ricci tensors are fixed by $[\\nabla_{a},\\nabla_{b}] u^{c} = -{{\\mathcal{R}}_{abd}}^{c} u^{d}$, and ${\\mathcal{R}}_{ab} \\equiv {{\\mathcal{R}}_{abc}}^{b}$. We use units with $c=G=1$ throughout. ", "conclusions": "Starting from a covariant and gauge-invariant formulation of the Boltzmann equation, we have derived a new expression for the CMB temperature anisotropy under the instantaneous recombination approximation, valid for scalar perturbations in open, closed and flat universes. Our expression uses only covariantly-defined variables, and is manifestly gauge-invariant. The result is more useful in multicomponent models with scalar perturbations than earlier covariant results~\\cite{ellis-er97,LC97-er,dunsby96b}. In the case of a universe which is baryon-dominated at recombination, we find a simple expression for the anisotropy which corrects a similar result by Dunsby~\\cite{dunsby96b}. By making use of numerical solutions to the perturbation equations, we have discussed the conditions on the last scattering surface and their contributions to the characteristic features of the CMB power spectrum. We ended with a discussion of the original Sachs-Wolfe calculation for the temperature anisotropy. We have discussed why it is not necessary to locate accurately the last scattering surface in such calculations (because of the compensation effect), and how the extra term in the Sachs-Wolfe calculation, reported recently by Ellis and Dunsby~\\cite{ellis-er97}, is missed in many calculations which employ a gauge-dependent ``expected temperature'', since the angular dependence of this temperature is often overlooked. For a universe which is matter dominated at recombination, but not necessarily adiabatic, the extra term is the spatial gradient of the radiation energy density in the energy-frame, $(\\bar{{\\mathcal{X}}}_{k}^{(\\gamma)} Q^{(k)} /4)_{A}$. For models with isothermal surfaces of last scattering, this inhomogeneity describes a distortion of the last scattering surface relative to the surfaces of simultaneity of the energy-frame. The extra redshift incurred by this distortion is a significant component of the temperature anisotropy on intermediate and small scales." }, "9804/astro-ph9804193_arXiv.txt": { "abstract": "We search OGLE-I photometric database for stars, which, as defined by formal criteria adopted by OGLE-I microlensing search, showed variability during only one out of 3 or 4 observing seasons. The results include 17 previously reported microlensing events, 2 newly discovered candidate events and 15 intrinsically variable stars that have a potential of contaminating samples of microlensing events. Based on photometry obtained in 1992 and 1993 OGLE \\#10 was tentatively included in the list of microlensing candidates, however its light curve in 1994 and 1995 shows many characteristics of the variable stars found in our search, and most likely it is not a microlensing event. For all stars which passed our tests, we provide $44 \\times 44$ arcsec ($101 \\times 101$ pixels) centered subframes from each OGLE-I frame in $I$ band. It is the first time when images used to derive photometry of microlensing events are available in convenient format to astronomical community. ", "introduction": "The search for rare cases of gravitational microlensing in the Local Group requires monitoring of $\\sim 10^6$ stars over several months in order to yield a significant rate of detections. A common implementation adopted is to construct a massive photometry database and subsequently select stars which experienced brightening of the type we expect on theoretical grounds (see Paczy\\'nski, 1996, for a recent review of basic theory, current microlensing searches and results). For the vast majority of events the light curves should follow a single point mass microlensing curve. Possible departures and exceptions from this basic case are extremely interesting. For random distribution of stars the probability that a given star will be lensed twice over duration of the experiment is negligible. Large fraction of binary lenses is expected to give raise to the population of ``wide binary'' light curves with two separate amplification regions ($\\sim 1\\%$ of the total rate of events). However in most cases the secondary peak should be weak ($A_{\\rm max} < 0.1$) and probably would only be discovered in light curves, which called attention because of the primary event (Di Stefano and Mao 1996). Photometry in very crowded fields, which are natural places to look for microlensing events, is often of limited quality. The majority of stars measured are just above the detection limit, many events have modest amplitudes, and numerous light curves are unevenly sampled. Therefore a microlensing curve may accidentally give a good fit to the brightening which is due to the intrinsic variability of the star. Certainly a repeated brightening of the same type would need to be very carefully examined before any claim of a detection of the wide binary microlensing repeater. As a result, it is generally hard to lower confusion rate without lowering the number of events returned by the procedure. For microlensing events reported by OGLE-I project (Udalski et al. 1992) the basic requirement was that variability should occur during only one observing season (Udalski et al. 1994a). It was assumed that a sample of stars selected according to the above condition contains the majority of microlensing events and relatively few variable stars of other types. In this paper we investigate how many variable stars have light curves that, given the time sampling of OGLE-I experiment, appear constant during all seasons except for just one. The quantitative information about stellar variability background, against which microlensing events are detected, allows further tuning of the methods used in automated detection of the events. ", "conclusions": "A general conclusion is that variability background in OGLE-I search was reasonably well separated from microlensing events, although OGLE \\#10, a single candidate which most likely is not a microlensing event, constitutes a 5\\% confusion rate. A variable star, just like any other star, may be amplified by microlensing, but an increase of brightness by 0.1 mag may be naturally explained by intrinsic variability of this star, especially that similar objects (certainly not microlensing events) apparently exist. We find two additional possible events that were not returned by the automated procedure of Udalski et al. (1994a). This is not surprising since we relax some of the selection cuts applied before. Moreover, one of those events happened near the end of the observing season while the other had very short time scale and poorly sampled light curve. Both of them are very inconspicuous. The outburst experienced by MM3 $I$ 58214 (most likely a flare or CV star) is an important case which has a potential of contaminating samples of microlensing candidates. Suppose we had no data just before the flare. With photometric accuracy comparable to OGLE-I data such variable could be taken for a fading microlensing event and uncertainty would have to be resolved by spectroscopy and/or monitoring of the star long after the event. We note a relatively large number of periodic or almost periodic variables which change amplitude. They mimic constant stars for time long enough to pass the most important criterion of the OGLE-I search, i.e., that a star should vary within a limited time interval with essentially constant flux at all other times. It is mostly due to relatively poor time coverage of the OGLE-I experiment and should not be difficult to overcome in the second phase of the project, OGLE-II (Udalski, Kubiak and Szyma\\'nski 1997). Two recommendations can be made for the future. First, low amplitude events, e.g. with $A_{\\rm max} < 1.5$, may be safely ignored in the calculation of the optical depth to prevent potential problems with objects similar to OGLE \\#10. MACHO team is already using such cut off to avoid contamination by ``bumpers''. Second, a requirement of roughly even photometric coverage of both sides of the magnification peak allows filtering out stars with (usually asymmetric) outbursts. OGLE-I events used in the optical depth determination satisfy such condition, however some of the events discovered by the EWS do not. The full set of data used in this paper, including $101 \\times 101$ pix subframes extracted from every $I$ band image of each object in Tables~1 and 2, is available for public. Images in FITS format as well as photometric data in standard Johnson system may be retrieved via anonymous ftp from {\\tt astro.princeton.edu (128.112.24.45)} -- directory {\\tt /ogle/var\\_background} and {\\tt sirius.astrouw.edu.pl (148.81.8.1)} -- directory /ogle/var\\_background. See {\\tt README} file for details." }, "9804/astro-ph9804299_arXiv.txt": { "abstract": "We present X-ray spectra obtained with \\sax\\ (Satellite per Astronomia X) of 10 BL Lac objects, selected from the Einstein Medium Sensitivity and Slew Surveys. We find that in about half of the objects a fit in the 0.1-10 keV range with a single power law and free absorption yields values of $N_{\\rm H}$ larger than the Galactic ones. In most of these cases, however, broken power law fits with $N_{\\rm H}$ fixed at the Galactic values yield an alternative, better description of the data and indicate a steepening of the spectrum with increasing energy. One object (1ES1101-232) is detected up to $\\sim$ 100 keV. Its spectral energy distribution (SED) peaks in the medium energy X-ray band. For each object we compute the peak frequency of the SED from multifrequency data. The spectral indices $\\alpha_x$ in the 2-10 keV band ($F_\\nu \\propto \\nu^{-\\alpha_x}$) are smaller (i.e. flatter spectrum) for objects with higher peak frequencies. We therefore confirm and extend to higher energies the behavior already known for X-ray selected BL Lac objects in the ROSAT band. We do not find spectral indices smaller than 1; however, the flat distribution of $\\alpha_x$ and the correlation between $\\alpha_x$ and peak frequency found from our data suggest that a number of objects may exist, which in the quiescent status have flatter spectrum and peak frequency in the hard X-ray range. ", "introduction": "\\noindent BL Lacertae objects are a rare type of Active Galactic Nuclei (AGN) characterized by strong and variable emission of non-thermal radiation across the entire electromagnetic spectrum, from radio waves to high energy $\\gamma$-rays. In three cases (Mkn 421: Punch et al. 1992; Mkn 501: Quinn et al. 1996; 1ES 2344+514: Catanese et al., 1998) the emission has been detected up to TeV energies. BL Lacertae objects comprise the most violent (highly and rapidly variable, highly polarized) and most elusive (extremely difficult to find in optical surveys) sources amongst AGN. Unlike most other AGN they do not show evidence (by definition) for strong emission lines or large Infra-Red or UV excesses. The emission from radio to $\\gamma$-rays can be explained as due to synchrotron radiation up to a certain maximum frequency (that ranges approximately from $10^{13}$ to $10^{17}$ Hz), above which a sharp turnover occurs until a second component due to Compton scattered radiation dominates, making these objects detectable up to the highest energies so far accessible (see e.g., Ulrich, Maraschi and Urry, 1997). The extreme properties of BL Lacs require that the matter emitting the radiation moves at relativistic speeds in the direction of the observer. The spectral change from synchrotron to Compton radiation is crucial for the understanding of the physics of BL Lacs. However, up to now this has been inferred only from the comparison of X-ray measurements carried out with different instruments and very often at different epochs. The wide energy band of \\sax\\ offers for the brightest objects the best opportunity to directly detect without ambiguity this spectral change and to study the X-ray spectra at the same epoch over a large interval. To this end we have undertaken a program that aims at studying in detail the X-ray spectrum of a large and well defined subsample of soft X-ray selected BL Lacs. This sample includes mostly objects that are expected to show strong spectral curvature and spectral breaks, since the synchrotron break should occur just before or in the \\sax\\ band. We aim at measuring in detail the shape of the most energetic part of the synchrotron emission, and trying to establish where and how the Compton component becomes dominant. We also intend to look for the correlation between spectral slope and break energy found in ROSAT data (Padovani $\\&$ Giommi, 1996; Lamer, Brunner \\& Staubert, 1996). ", "conclusions": "We have analyzed the spectra for 10 X-ray selected BL Lacs observed with the Narrow Field Instruments on board the \\sax\\ satellite. The sources are detected from $\\sim 0.2$ up to $\\sim$ 10 keV (and in one case up to $\\sim 100$ keV with the PDS instrument) with a very smooth appearance. The spectrum is generally well fitted by either a single power law, or by a broken convex power law that most probably represents the steepening after the synchrotron peak, whose position is determined also by using simultaneous optical observations. Variability is not present during the short \\sax\\ exposure; analysis of ROSAT data shows for most of the sources little variability (within $\\sim 30\\%$) with respect to the \\sax\\ flux and spectral indices consistent with the \\sax\\ ones. The spectral energy distributions, which include literature data, instead show variability in all bands. The X-ray spectral indices $\\alpha_x$ range between 1 and 1.5 with a flat distribution and a mean value $\\langle \\alpha_x \\rangle = 1.31\\pm0.06$. The scatter in the distribution is due to an anti-correlation we have found between $\\alpha_x$ and the frequency of the peak of the emission, $\\nu_{peak}$. This extends to the \\sax\\ band a correlation which had been discovered in the ROSAT band for this class of objects. The fact that sources with harder X-ray spectra have higher $\\nu_{peak}$ is expected if the \\sax\\ band is still dominated by synchrotron emission, which is also consistent with the spectral energy distributions of our BL Lacs. Furthermore, we have no evidence of a spectral flattening (indicating the arising of the Compton component) in the present spectra, but future PDS detections, that are possible with exposure times slightly longer than those obtained here, might help in this respect. The large fraction (at least 2 out of 10) of HBL selected in the soft X-ray band found with a flat ($\\alpha_x \\sim 1$) X-ray slope (i.e., they are near the peak of the synchrotron emission) and the distribution of $\\alpha_x$ values support the view that objects with even higher spectral peaks in their quiescent status indeed exist, and might be found in large numbers if we devise the correct strategy (e.g., samples at harder X-rays, TeV sources, etc.) Moreover, these sources are good candidates to be TeV {\\it emitters}. In fact, in the sources with the flattest $\\alpha_x$ the peak of the synchrotron component is localized in the soft X--ray range. Electrons emitting at 1 keV by the synchrotron process have Lorentz factors $\\gamma\\sim 2.5\\times 10^5 (\\nu_{peak,1~keV}/B\\delta)^{1/2}$, where $\\nu_{peak} = 2.42 \\times 10^{17} \\nu_{peak,1~keV}$ Hz, $B$ is the value of the magnetic field in Gauss and $\\delta$ is the usual Doppler factor. Through the inverse Compton mechanism, they can emit up to $E\\sim \\gamma m_ec^2\\delta\\sim 130 (\\nu_{peak,1~keV}\\delta/B)^{1/2}$ GeV. If the magnetic and radiation energy densities are equal (as it is, approximately, in the three BL Lacs already detected in the TeV band), the flux level of the synchrotron and inverse Compton peaks is roughly equal. Low redshift sources are therefore good candidates to be {\\it detected} in the GeV--TeV band, while the high energy emission of the more distant sources could be absorbed in $\\gamma$--$\\gamma$ interactions with the background IR photon field, whose intensity is still uncertain. Indeed, a cutoff in the high energy spectrum could be used to determine the IR background (see, e.g. Stecker \\& De Jager, 1997). This \\sax\\ project is still ongoing. We expect therefore to increase considerably the sample of soft X-ray selected BL Lacs for which we measure the spectrum in the 0.2-10 keV range and possibly above. With a larger complete sample, and combining the results with other complementary \\sax\\ projects, we expect to be able to draw a clearer picture of the relationship between the local X-ray slope and the overall energy distribution of this class of sources, in order to derive firmer conclusions on the behaviour at hard X-ray energies and on the mechanisms of the emission." }, "9804/astro-ph9804066_arXiv.txt": { "abstract": "About 25\\% of the optical extragalactic sky is obscured by the dust and stars of our Milky Way. Dynamically important structures might still lie hidden in this zone. Various approaches are presently being employed to uncover the galaxy distribution in this Zone of Avoidance (ZOA). Results as well as the different limitations and selection effects from these multi-wavelengths explorations are being discussed. Galaxies within the innermost part of the Milky Way --- typically at a foreground obscuration in the blue of $A_{\\rm B} \\ga 5^m$ and $|b| \\la \\pm5\\degr$ --- remain particularly difficult to uncover except for H\\,{\\sc i}-surveys: the Galaxy is fully transparent at the 21cm line and H\\,{\\sc i}-rich galaxies are easy to trace. We will report here on the first results from the systematic blind H\\,{\\sc i}-search ($v \\leq 12700$ km\\,s$^{-1}$) in the southern Zone of Avoidance which is currently being conducted with the Parkes Multibeam (MB) Receiver. ", "introduction": "To understand the dynamics within the local Universe -- the mass distribution and the local velocity field with its peculiar and streaming motions -- a detailed map of the 3-dimensional galaxy distribution is highly desirable. However, the dust extinction and confusion with stars in the disk of our Galaxy make this very difficult for $\\sim$25\\% of the sky, and the following questions remain unanswered: Could a nearby Andromeda-like galaxy have escaped detection to date, hence change our understanding of the internal dynamics and mass derivations of the Local Group (LG), and the present density of the Universe from timing arguments (Peebles 1994)? Is the dipole in the Cosmic Microwave Background Radiation (direction and amplitude) entirely explained by the gravity on the LG from the irregular mass/galaxy distribution? As the nearest galaxies ($v<300$ \\kms) generate 20\\% of the total dipole moment (Kraan-Korteweg 1989) nearby individual galaxies are equally important as massive groups, clusters and voids. Is the mass overdensity in the Great Attractor (GA) region -- postulated from a large-scale systematic flow of galaxies towards ($\\ell,b,v)\\sim(320\\degr,0\\degr,4500$\\kms) (Kolatt \\etal\\ 1995) -- in the form of galaxies, hence does light trace mass? Does the Supergalactic Plane, other superclusters, walls and voids connect across the Milky Way and might other large-scale structures (LSS) have gone undetected due to this 'zone of avoidance'? ", "conclusions": "The combination of the complementary multiwavelength surveys allow a new probing of LSS in the 'former' ZOA. The \\HI\\ surveys are particularly powerful at the lowest latitudes. But future merging of ZOA data with catalogs outside the ZOA will have to be done with care to obtain 'unbiased' whole-sky surveys. From the sensitivity attained with the first 2 scans of the ZOA MB-survey it can be maintained that no Andromeda or other \\HI-rich Circinus galaxy is lurking undetected behind the extinction layer of the southern Milky Way.\\\\ {\\sl Acknowledgements} --- The help of the HIPASS ZOA team members R.D. Ekers, A.J. Green, R.F. Haynes, P.A. Henning, R.M. Price, E. Sadler, and L. Staveley-Smith is gratefully acknowledged." }, "9804/astro-ph9804072_arXiv.txt": { "abstract": "Multicolour images of the starbursting metal poor blue compact galaxy ESO~338-IG04 have been obtained with the {\\it Wide Field Planetary Camera 2} on board the {\\it Hubble Space Telescope}. In the images we find numerous point-like sources concentrated towards the main body of the galaxy, which we identify as globular cluster candidates. We show that these objects are physically associated with the galaxy and that they are spatially extended. Given their high intrinsic luminosities, these objects cannot be individual stars. Using photometric evolution models we show that the objects constitute a rich population of massive star clusters with ages ranging from a few Myr to $\\sim$ 10 Gyr, and masses ranging from $10^4$ to more than $10^7 {\\cal M_{\\odot}}$. There are peaks in the age distribution of the clusters: one with objects $\\le30$ Myr, one at $\\sim 100$~ Myr, one at $\\sim 600$~ Myr, one to two at $2.5-5$ Gyr and one at $\\sim10$ Gyr. The youngest objects are predominantly found in the crowded starburst region. They have properties which agree with what is expected for young globular clusters, although it cannot be excluded that some of them may be dissolved or disrupted. For objects older than a few times 10 Myr, the only plausible explanation is that these are globular clusters. The galaxy presently appears to be involved in a merger, which is the probable cause of the present globular cluster formation. The presence of a numerous intermediate age (2.5 to 5 Gyr) population of globular clusters, suggests that a previous merger might have occurred. As the starburst fades, this galaxy will become very rich in globular clusters. Transforming all objects to an age comparable to that of Milky Way globular clusters reveals a luminosity function similar to the Galactic. We suggest that this galaxy is the result of a merger between a dwarf elliptical and a gas rich dwarf. The possibility of dating the globular clusters offers an efficient way of studying the history of violent star formation in this and similar galaxies. ", "introduction": "Globular clusters (GCs) are generally old stellar systems and are believed to be the first objects to form in the process of the formation of a galaxy. This is supported by age estimates from observations of GCs in the Milky Way and other nearby galaxies. The Large Magellanic Cloud (LMC) is however known to host several young blue \"populous\" clusters, which could be young globular clusters. In the recent years blue globular cluster candidates have been found in some interacting/merging galaxies such as NGC~3597 (Lutz 1990, Holtzman et al. 1996), NGC~7252 (Whitmore et al. 1993), \"The Antennae\" (Whitmore \\& Schweizer 1995) and NGC~3921 (Schweizer et al. 1996). This has strengthened the idea that GCs can form not only when galaxies form, but also when galaxies are reformed in the process of galaxy mergers (Schweizer 1986, Ashman \\& Zepf 1992, Whitmore 1996). This could possibly also circumvent the problem, faced by the idea that elliptical galaxies form by the merging of late type galaxies, that ellipticals have higher specific frequencies of globular clusters (van den Bergh 1984, 1994). Understanding how and when GCs can form will thus not only help us in understanding these systems, but also aid in understanding galaxy formation and evolution. Studies of actively star-forming galaxies, e.g. Henize 2-10 (Conti et al. 1994), NGC~1244 (Hunter et al. 1994), NGC~1705 (O'Connell et al. 1994), NGC~1569 (De Marchi et al. 1997) and M82 (O'Connell 1995) have revealed the presence of star clusters with luminosities comparable to R126, the central cluster of 30 Doradus in the LMC, and higher. Meurer et al.(1995) used the HST for an ultraviolet study of nine starburst galaxies and found bright \"super star clusters\" in all. These clusters are generally bluer than similar objects found in interacting galaxies, and are preferentially found in the central regions of the starbursts. It has been proposed (e.g. Conti et al. 1994) that these blue objects might be forming globular clusters. This is however still an open question since it is not clear if these systems will survive as gravitationally bound systems. Uncertainties in the value of the (universal ?) stellar initial mass function (IMF) makes the mass estimates, mainly based on ultraviolet data, highly uncertain, especially since some studies only includes one spectral bandpass. Many of these galaxies appear to be involved in some form of interaction. One could interpret these observations as evidence that massive star clusters can form in regions of very active star formation, such as giant extragalactic HII-regions (GEHRs) like 30 Doradus. Kennicutt and Chu (1988) made a statistical investigation of data on extragalactic blue populous clusters and giant HII-regions. They concluded that the young clusters in LMC are not luminous enough to evolve into globular clusters comparable to the massive galactic ones, and that far from all GEHRs will produce young populous clusters that could become GCs. Blue compact galaxies (BCGs) are characterised by their blue colours ($B-V \\le 0.5$), strong nebular emission lines, indicative of the formation of relatively hot (massive) stars, and low chemical abundances. The derived star formation rates could, considering the gas supply, only be sustained for a small fraction of a Hubble time. This together with the low metallicities (with IZw18 being the extreme in this sense) once lead to the idea that BCGs might be genuinely young objects now experiencing their first star formation epoch (Sargent and Searle 1970). Now, most BCGs are believed to be old, experiencing recurrent bursts of star formation intervened by long quiescent periods. Still we do not yet understand what triggers these bursts of star formation. One possible trigger mechanism is tidal or direct interactions with companion galaxies or gas clouds. In this paper we will present conclusive evidence for young and old globular clusters in the blue compact galaxy ESO~338-IG04. Section 2 describes the observations and Sect. 3 describes how photometry was performed on the globular cluster candidates. In Sect. 4 we show that the objects are physically associated with the galaxy and that they are spatially resolved. In Sect. 5 we discuss how the observed photometric properties can be interpreted in terms of age and mass of the objects. Section 6 includes a further discussion on the nature of the cluster candidates, and Sect. 7 contains the conclusions. \\subsection{General properties of the target galaxy} ESO~338-IG04, also known as Tol~1924-416, resides at a distance of 37.5 Mpc ($v_{hel}=2813$ km~s$^{-1}$, $H_0= 75$ km s$^{-1}$ Mpc$^{-1}$; this value will be used throughout the rest of the paper) at the celestial coordinates $\\alpha_{1950}=19^{\\rm h}~24^{\\rm m}~30^{\\rm s}~ \\delta_{1950} = -41\\degr~34\\arcmin~00\\arcsec$. It is intrinsically bright ($M_V=-19.3$) and blue ($B-V=0.4$), (Bergvall and \\\"Ostlin 1998). The oxygen abundance is $12 \\%$ of the solar value (Bergvall, 1985). It's physical size is small, $8.5 \\times 4.5$ kpc measured at $\\mu _V=25 {\\rm mag}~{\\rm arcsec}^{-2}$, the non dwarfish luminosity being due to the active star burst. It has an integrated HI-mass of $3\\times10^9 {\\cal M_{\\odot}}$ (\\\"Ostlin et al. 1997b). The optical velocity field of ESO~338-IG04 show rotation aligned with it's apparent major axis, though with several large scale irregular features, and it is most easily understood as a merger between two galaxies or a galaxy and a gas cloud (\\\"Ostlin et al. 1997a). This is supported by the optical tail, extending towards the east. It has a spectroscopically confirmed companion galaxy (Bergvall unpublished; \\\"Ostlin et al. 1997a) which lies at a projected distance of 70 kpc, see Fig. 3. The companion is a somewhat fainter ($M_V = -17.9$), star forming galaxy which shows regular rotation (\\\"Ostlin et al. 1997a). Bergvall noticed, in ground based images, an apparent concentration of faint blobs around the galaxy, which he interpreted as globular cluster candidates. We have re-observed this galaxy with the HST and will in the subsequent sections show that there is substantial evidence that this galaxy has formed massive globular clusters at several epochs. \\begin{figure} \\picplace{8.5cm} \\caption[]{A Digitized Sky Survey (DSS) image (obtained with $Skyview$) showing the target galaxy and its companion. North is up and east is left. The total angular size of the field is ~$9\\arcmin \\times 9\\arcmin$. ESO~338-IG04 is at the upper left, the companion at the lower right. } \\end{figure} In July 1997, we obtained spectra of a few of the brightest globular cluster candidates, using the ESO New Technology Telescope (NTT), at La Silla, Chile. The results from this spectroscopic study will be presented in a future publication. ", "conclusions": "Multi-colour photometry with HST/WFPC2 of the metal poor blue compact galaxy ESO~338-IG04 (Tol~1924-416) has revealed a rich population of faint point-like sources in and surrounding the galaxy. Aperture photometry has been performed on these, and the photometric results have been transformed into ages and masses using a spectral evolutionary synthesis model. Special care was taken to assure that the sources are physically associated and not chance projected background or foreground sources. The results can be summarised as follows: \\begin{enumerate} \\item The objects discussed in this paper have absolute magnitudes, ~$M_v$, in the range -7 to -15 and $(v-i)$ colours ranging from less than -1 to almost 2. \\item The vast majority of the found objects are spatially resolved star clusters, physically associated with the galaxy. The outer objects follow the luminosity distribution of the galaxy closely (Fig. 9). \\item The number of interlopers, i.e. foreground stars, background galaxies and super giants in the target galaxy, in the presented sample is conservatively estimated to be less than ten objects. This leaves us with more than 112 detected star clusters in ESO~338-IG04. \\item Using the photometric evolution model we show that the objects, in general, are well fitted by Salpeter and Miller-Scalo IMFs, and that the resulting age distribution is insensitive to the adopted IMF. The ages of the clusters range from a few Myr to more than 10 Gyr. \\item The objects have inferred masses ranging from $10^4$ to more than $10^7 {\\cal M_{\\odot}}$. \\item The above listed properties show that the objects are massive globular clusters of varying age. \\item There are several peaks in the globular cluster age distribution. This shows that there have been several globular cluster forming events in this galaxy. These appear to have occurred ~$\\sim$ 11 Gyr ago, 2.5-5 Gyr ago, 600 Myr ago, 100 Myr ago and \"now\". The present event has lasted for a few times 10 Myrs. \\item More than half of the total expected number of GCs have an age in the range 2 to 5 Gyr. This peak consists of two subpopulations with either different age ($\\Delta_{\\rm AGE} \\sim 2.5$~Gyr) or metallicity ($\\Delta_{\\rm [Fe/H]} \\sim 1$~dex). \\item The specific frequency, $S_N$, is high in this galaxy. Taking into account that the fading starburst will decrease the luminosity of the galaxy by 1.5 magnitudes in 1 Gyr, this galaxy will be come very rich in globular clusters, even if the newly formed GCs don't survive. The predicted specific frequency of globular clusters is much larger than for late type galaxies and comparable to that for giant ellipticals. \\item We suggest that a merger is responsible for the current starburst and cluster formation. In view of the high specific frequency of old globular clusters suggests that an elliptical galaxy is a main ingredient in this system. To provide gas for star and cluster formation a gas rich dwarf must be the other main ingredient. A major formation event 2.5-5 Gyr ago suggests that a merger involving one gas rich component also occurred at that time. \\item In view of the high specific frequency and the present merger, we suggest that this galaxy will evolve into a moderately luminous elliptical galaxy, unless disturbed by possible future interactions. \\item This investigation for the first time gives strong support for a rich population both old and newly formed GCs in a metal poor blue compact galaxy. \\item Detection and dating of GCs in moderately distant BCGs is feasible using the HST and offers an efficient way of studying the violent star formation history, if also the effects of metallicity and extinction can be handled. \\end{enumerate}" }, "9804/astro-ph9804244_arXiv.txt": { "abstract": "Recent work suggests that rich clusters of galaxies commonly have large populations of dwarf (ie. low luminosity) members, that is their luminosity function (LF) turns up to a steep slope at the faint end. This population, or more particularly the relative numbers of dwarfs to giants, appears to be very similar for clusters of similar morphology, but may vary between cluster types. We have previously suggested that dwarfs may be more common in less compact, spiral rich clusters. Similarly we have found evidence for population gradients across clusters, in that the dwarf population appears more spatially extended. In the present paper we summarise the current evidence and propose, in analogy to the well-known morphology - density relation, that what we are seeing is a dwarf population - density relation; dwarfs are more common in lower density environments. Finally we discuss recent semi-analytic models of galaxy formation in the hierarchical clustering picture, which may give clues as to the origin of our proposed relation. ", "introduction": "Much recent work has been devoted to the question of the galaxy luminosity function (LF) within rich clusters, particularly with regard to the faint end which has become accessible to detailed study through various technical and observational improvements (see e.g., Driver et al. 1994; Biviano et al. 1995; Bernstein et al. 1995; Mohr et al. 1996; Wilson et al. 1997; Smith, Driver \\& Phillipps 1997 = Paper I; Trentham 1997a,b). For the most part these studies concur that the LF becomes steep (Schechter (1976) slope $\\alpha \\leq - 1.5$) faintwards of about $M_{B} = -17.5$ or $M_{R} \\simeq -19$ (for $H_{0}$ = 50 km s$^{-1}$ Mpc$^{-1}$), and Paper I suggested that such a dwarf rich population might be ubiquitous. In a subsequent paper (Driver, Couch \\& Phillipps 1997a = Paper III) we have examined the luminosity distribution in and across a variety of clusters, examining the possible dependence of the dwarf population (in particular the ratio of dwarfs to giants) on cluster type and position within the cluster. In the present paper we summarise the evidence to date for the (dis)similarity of the dwarf population in different environments. ", "conclusions": "As with the corresponding morphology density relation for giant galaxies, the cause of our population - density relation could be either `nature' or `nurture', ie. initial conditions or evolution. Some clues may be provided by the most recent semi-analytic models of galaxy formation, which have been able to account in a general way for the excess of (giant) early type galaxies in dense environments (e.g., Baugh, Cole \\& Frenk 1996). The steep faint end slope of the LF appears to be a generic result of hierarchical clustering models \\footnote{ And was considered a problem until observational evidence for steep LFs increased!} (e.g., White \\& Frenk 1991; Frenk et al. 1996; Kauffmann, Nusser \\& Steinmetz 1997), so is naturally accounted for in the current generation of models. The general hierarchical formation picture envisages (mainly baryonic) galaxies forming at the cores of dark matter halos. The halos themselves merge according to the general Press-Schechter (1974) prescription to generate the present day halo mass function. However the galaxies can retain their individual identities within the growing dark halos, because of their much longer merging time scales. The accretion of small halos by a large one then results in the main galaxy (or cluster of galaxies, for very large mass halos) acquiring a number of smaller satellites (or the cluster gaining additional, less tightly bound, members). Kauffmann et al. (1997) have presented a detailed study of the distribution of the luminosities of galaxies expected to be associated with a single halo of given mass. The LFs are somewhat disjoint owing to the specific halo masses modelled; especially for the low mass halo there is a preferred luminosity for the central galaxy plus a tail to lower luminosities. For a realistic mix of halo masses, these would no doubt be smoothed to look more like conventionally observed LFs. Nevertheless, we can still easily compare the numbers of dwarf galaxies per unit giant galaxy luminosity (rather than the amplitude of the giants' LF) between halos of different mass. The Kauffmann et al. models mimic a \"Milky Way system\" (halo mass $5 \\times 10^{12} M_{\\odot}$), a sizeable group (halo mass $5 \\times 10^{13} M_{\\odot}$) and a cluster mass halo ($10^{15} M_{\\odot}$). Using their figure 2 (which also emphasises the identical faint end slopes predicted for all the different environments), we choose to quantify the number of dwarfs by $N_{-18}$, the number of dwarfs per system in the $M_{B} = -18$ bin. Because of the very similar slopes, the choice of bin or range of bins does not affect our conclusions, so this is the equivalent of the total number of dwarfs used in Figure 1. To quantify the giant population we choose the total light of galaxies of $M_{B} = -20$ or brighter, in units of $L_{*}$ galaxies (taking $M_{B}^{*} = -21$), which we call $N_{-21}$. Using this definition, rather than the actual number of galaxies brighter than some value (as in our observational data) allows for the discretization of the LFs for small halos. The results are summarized in Table 1. The ratio of these two values $N_{-18}$ and $N_{-21}$ then quantifies the relative dwarf galaxy populations. Roughly speaking, for smooth LFs with a shape similar to that observed, we should multiply these values by about 5, giving a range from about 1 to 3, to compare with our observational DGRs. \\begin{table*} \\begin{center} \\begin{tabular} {lccccccccr} Halo Mass ($M_{\\odot}$) & $N_{-18}$ & $N_{-21}$ & $N_{-18}/N_{-21}$ \\\\ \\tableline $5 \\times 10^{12}$ & 0.2 & 0.34 & 0.6 \\\\ $5 \\times 10^{13}$ & 2.2 & 3.8 & 0.6 \\\\ $1 \\times 10^{15}$ & 40 & 190 & 0.2\\\\ \\end{tabular} \\end{center} \\caption{Dwarf numbers as a function of halo mass. \\label{tbl-1}} \\end{table*} We see that the Milky Way and small group halos have similar numbers of dwarf galaxies per unit giant galaxy light, whereas the dense cluster environment has a much smaller number of dwarfs for a given total giant galaxy luminosity. Thus the predictions of the hierarchical models (which depend, of course, on the merger history of the galaxies) are in general agreement with our empirical results if we identify loose clusters and the outskirts of rich clusters with a population of (infalling?) groups (cf. Abraham et al. 1996), whereas the central dense regions of the clusters originate from already massive dark halos. By inputting realistic star formation laws etc., Kauffmann et al. can further identify the galaxies in the most massive halos with old elliptical galaxies, and those in low mass halos with galaxies with continued star formation. This would imply the likelihood that our dwarfs in low density regions may still be star forming, or at least have had star formation in the relatively recent past (cf. Phillipps \\& Driver 1995 and references therein). Note, too, that these galaxy formation models would also indicate that the usual (giant) morphology - density relation and our (dwarf) population - density relation arise in basically the same way. Finally, we can see that if these models are reasonably believable, then we need not expect the field to be even richer in dwarfs than loose clusters; the dwarf to giant ratio seems to level off at the densities reached in fairly large groups. To summarise, then, we suggest that the current data on the relative numbers of dwarf galaxies in different clusters and groups can be understood in terms of a general dwarf population versus local galaxy density relation, similar to the well known morphology - density relation for giants. Low density environments are the preferred habitat of low luminosity galaxies; in dense regions they occur in similar numbers to giants, but at low densities dwarfs dominate numerically by a large factor. This fits in with the general idea that low luminosity galaxies are less clustered than high luminosity ones (particularly giant ellipticals). Plausible theoretical justifications for the population - density relation can be found within the context of current semi-analytic models of hierarchical structure formation." }, "9804/gr-qc9804086_arXiv.txt": { "abstract": "Inflation of cosmic gauge and global strings is investigated by numerically solving the combined Einstein and field equations. Above some critical symmetry-breaking scales, the strings undergo inflation along the radial direction as well as the axial direction at the core. The nonsingular nature of the spacetimes around supercritical gauge and global strings is discussed and contrasted to the singular static solutions that have been discussed in the literature. ", "introduction": "Cosmic strings are linelike topological defects that may form as a result of a phase transition in the early universe. If a string is associated with a magnetic field, it is called a gauge string, otherwise it is a global string. They have attracted much attention because of their cosmological importance: deficit angle in the spacetime geometry and a candidate for the seed of structure formation in the early universe. It was proposed that topological defects can inflate if the symmetry-breaking scale satisfies $\\eta \\gtrsim \\eta_c \\sim {\\cal O}(m_p)$ in Refs.~\\cite{Linde,Vilenkin}. This was later verified in numerical simulations by Sakai {\\it et al.}~\\cite{Sakai}. They found, in particular, that the critical value of $\\eta$ for domain walls and global monopoles is $\\eta_c \\simeq 0.33m_p$. Then what about cosmic strings? There is no reason why we exclude cosmic strings out of the topological inflationary category. Recently, it was numerically proved that a (2+1) dimensional gauge string can inflate by de Laix {\\it et al.}~\\cite{Tanmay}. In this paper, we shall numerically solve the combined Einstein and field equations for a gauge and a global string in (3+1) spacetime dimensions. For the gauge string, we find that the core inflates if $\\eta \\gtrsim 0.25m_p$ with unit winding number in the critical coupling case (Bogomol'nyi limit). For the global string, $\\eta \\gtrsim 0.23m_p$. The critical values decrease as the winding number increases. For the gauge string, the critical value also decreases slightly as the coupling of the gauge field to the scalar field becomes weaker than the self coupling of the scalar field. The asymptotic spacetime of a gauge string is known to be conical~\\cite{VilenkinC}. This spacetime exhibits a deficit angle $\\Delta =8\\pi G\\mu$, where $\\mu\\sim \\eta^2$ is the mass per unit length of the string. When the symmetry-breaking scale is sufficiently large, the deficit angle exceeds $2\\pi$ and analyses of the static solution show that the spacetime possesses a physical singularity outside the core of the string~\\cite{Gott,Ortiz,Laguna}. However, from our numerical results we know that supermassive strings are dynamical and undergo inflation at the core. Therefore, we believe that the static treatment of supermassive strings loses its validity and that we should treat them in a time-dependent way. For global strings, the singularity exists regardless of the symmetry-breaking scale. Many people have tried to find a static solution of a global string and they found that there also exists a physical singularity outside the core of the string~\\cite{Cohen,Sikivie,Gregorys,Gibbons}. What was suggested to remove this singularity is again a time-dependent treatment of the string. Gregory~\\cite{Gregoryt} introduced a specific metric which has an axial time dependence and showed that this spacetime is nonsingular. In our work, we follow the evolution of supermassive gauge and global strings in a general time-dependent metric and show that no singularity develops in the spacetimes around the strings. In the next section, we solve the Abelian Higgs model of a gauge string and discuss its inflation and spacetime geometry. Sec.~III is devoted to global strings. Our conclusions are summarized in Sec.~IV. In Appendix, we show the equations in detail and the numerical algorithms. ", "conclusions": "We have investigated inflation in cosmic strings. In the core region, the strings undergo inflation radially as well as axially when $\\eta \\gtrsim \\eta_c$. With unit winding number ($n=1$) the critical values for inflation were found to be $\\eta_c \\approx 0.25m_p$ for a gauge string in the Bogomol'nyi limit ($\\beta=1$) and $\\eta_c =0.23m_p$ for a global string. The critical values decrease as $n$ and $\\beta$ increase. We have explained this $\\eta_c$ variation in terms of the core size of defects. The core of defects inflates when its size becomes bigger than the horizon scale: for larger $n$ and $\\beta$, strings have bigger cores, and the global string has a bigger core than the gauge string for a given $\\eta$. Regardless of the symmetry-breaking scale $\\eta$, around the center of defects the de Sitter expansion is established since the scalar field stays about the top of the potenital ($\\phi\\approx 0$). However, this is not sufficient for the cores of defects to inflate. Inflation requires another condition which is the core size being comparable to the horizon scale so that the core can be dynamical due to the gravitational effect. Or equivalently, the potential $V(\\phi)$ needs to be flat enough at $\\phi\\approx 0$ so that the field $\\phi$ can spend enough time about the top of the potential. For this condition to be satisfied, the symmetry-breaking scale $\\eta$ needs to be sufficiently large. This description also explains why we have somewhat lower critical values of $\\eta$ for strings than those for domain walls and global monopoles ($\\eta_c\\approx 0.33m_p$). Strings have bigger cores at the same symmetry-breaking scale than the other defects. For supermassive gauge strings and all scale global strings, we have had troublesome physical singularities outside the core when we treat them in a static way. The elegant exit to nonsingular spacetimes is to introduce a time-dependent treatment. From the numerical simulations we could show that there is no singularity developing around time-dependent supermassive strings." }, "9804/astro-ph9804008_arXiv.txt": { "abstract": "}[1]{{\\footnotesize \\noindent {\\bf Abstract} #1 \\\\}} \\renewcommand{\\author}[1]{\\subsubsection*{#1}} \\newcommand{\\address}[1]{\\subsubsection*{\\it#1}} \\setlength{\\textheight}{20cm} \\setlength{\\textwidth}{13.5cm} \\begin{document} \\chapter*{Galactic Environments of the Sun and Cool Stars\\footnote{To be published in {\\it Planetary Systems -- The Long View}, eds. L. M. Celnikier and J. Tran Than Van, Editions Frontieres, 1998}} \\author{Priscilla C. Frisch} \\address{University of Chicago, Dept. Astronomy and Astrophysics, 5640 S. Ellis Ave., Chicago, IL 60637} \\abstract{ The importance of understanding the current and historical galactic environments of cool stars is discussed. The penetration of interstellar gas into a stellar astrosphere is a function of the interaction of the star with the interstellar cloud surrounding the star, and this factor needs to be understood if an efficient search for life-bearing planets is to be made. For the Sun, both current and historical galactic conditions are such that if a solar wind were present, it would have excluded most inflowing interstellar matter from the inner regions of the heliosphere for the past few million years. Variations in heliosphere size over the recent historical path of the Sun are estimated, along with estimates of astrosphere sizes for selected nearby stars. Considering only possible effects due to encounters with interstellar clouds, stable planetary climates are more likely for inner than outer planets.} ", "introduction": "The Sun moves through space at a velocity of about 17 pc per million years. This motion, combined with interstellar cloud motions driven by stellar evolution, yield a constantly changing galactic environment for the Sun and solar system. This environment affects the interplanetary environments of both outer and inner planets in the solar system, including Sun--Earth coupling mechanisms. By analogy, the interactions between other cool stars and the galactic environment of that star needs to be understood as part of the process of identifying planets conducive to ``higher'' life forms. The interstellar cloud surrounding the Sun at this time (known as the ``local interstellar cloud'', LIC), is warm, low density, and partially ionized: T$\\approx$7,000 K, n(H$^{\\rm \\circ})$$\\approx$0.2 cm$^{- 3}$, and n(e$^{-}$)$\\approx$0.1 cm$^{-3}$. The standard assumption for diffuse interstellar clouds is that n(p$^{+}$)=n(e$^{-}$). On the scale of typical cloud densities, the LIC is rather tenuous, and notably lower density than the 1 au solar wind density (see Fig. \\ref{densities}). This accounts for the ability of the solar wind today to exclude most interstellar material from 1 au. \\begin{figure}[ht] \\vspace*{4in} \\special{psfile=density.eps voffset=-107 vscale=60 hscale=60 } \\caption{Typical densities for material in our Galaxy. \\label{densities}} \\end{figure} In this paper the basis for understanding the relation between the properties of stellar wind envelopes around cool star systems and the physical properties of the surrounding interstellar clouds are examined. The author believes that the historical galactic environment of a star would have a direct impact on the stability of planetary atmospheres, and therefore on the distribution of intelligent life forms. This conclusion rests partly on the observation that the solar system has been in a region of space virtually devoid of interstellar matter over the past several million years \\cite{fy86}, \\cite{journey}. Additional reviews on interstellar matter (ISM) within the solar system can be found in the book {\\it The Heliosphere in the Local Interstellar Medium} \\cite{rudi}. For more information on the properties of local ISM (LISM) see \\cite{fr95}. For more information on nearby G-star space motions and environments, and the use of astrospheres as a test for interstellar pressure, see \\cite{fr93}. ", "conclusions": "" }, "9804/astro-ph9804187_arXiv.txt": { "abstract": "We evaluate in a homogeneous way the optical masses of 170 nearby clusters ($z\\le 0.15$). The sample includes both data from the literature and the new ENACS data (Katgert et al. 1996, 1998). On the assumption that mass follows the galaxy distribution, we compute the masses of each cluster by applying the virial theorem to the member galaxies. We constrain the masses of very substructured clusters (about $10\\%$ of our clusters) between two limiting values. After appropriate rescaling to the X-ray radii, we compare our optical mass estimates to those derived from X-ray analyses, which we compiled from the literature (for 66 clusters). We find a good overall agreement. This agreement is expected in the framework of two common assumptions: that mass follows the galaxy distribution, and that clusters are not far from a situation of dynamical equilibrium with both gas and galaxies reflecting the same underlying mass distribution. We stress that our study strongly supports the reliability of present cluster mass estimates derived from X-ray analyses and/or (appropriate) optical analyses. \\vspace*{6pt} \\noindent {\\em Subject headings: } galaxies: clusters: general - galaxies: distances and redshifts - X-rays: galaxies - cosmology: observations. ", "introduction": "The knowledge of the properties of galaxy clusters plays an important role in the study of large scale structure formation. In particular, the observational distribution of the abundance of galaxy clusters as a function of their mass places a strong constraint on cosmological models (e.g., Bahcall \\& Cen 1993; Borgani et al. 1997; Gross et al. 1998; White, Efstathiou \\& Frenk 1993). Moreover, recent studies stress the need for having reliable estimates of cluster masses to constrain the ratio between the baryonic to total mass and the consequent value of $\\Omega_0$ (e.g., White \\& Frenk 1991; White et al. 1993b). Indeed, the estimate of cluster masses is not an easy task, in spite of the various methods which are available. The application of the virial theorem to positions and velocities of cluster member galaxies is the oldest method of cluster mass determination (e.g., Zwicky 1933). More recent methods are based on the dynamical analysis of hot X-ray emitting gas (e.g., Cowie, Henriksen, \\& Mushotzky 1987; Eyles et al. 1991) and on gravitational lensing of background galaxies (e.g Grossman \\& Narayan 1989). Mass estimates derived from the dynamical analysis of gas or member galaxies which are based on the Jeans equation or its derivations, such as the virial theorem, assume that clusters are systems in dynamical equilibrium (e.g., Binney \\& Tremaine 1987). This assumption is not strictly valid; in fact, although clusters are bound galaxy systems, they have collapsed very recently or are just now collapsing, as is suggested by the frequent presence of substructures (e.g., West 1994). However, some analyses suggest that the estimate of optical virial mass is robust against the presence of small substructures (Escalera et al. 1994; Girardi et al. 1997a; see also Bird 1995 for a partially different result), although it is affected by strong substructures (e.g., Pinkney et al. 1996). Similar results come from studies based on numerical simulation (e.g., Schindler 1996a; Evrard, Metzler, \\& Navarro 1996; Roettiger, Burns, \\& Loken 1996) for X-ray masses estimated with the standard $\\beta$-model approach (Cavaliere \\& Fusco Femiano 1976), although some authors have claimed there is a systematic mass underestimation (e.g.; Bartelmann \\& Steinmetz 1996). Dynamical analyses based on galaxies have the further drawback that the mass distribution or (alternatively) the velocity anisotropy of galaxy orbits should be known a priori. Unfortunately, the two quantities cannot be disentangled in the analysis of the observed velocity dispersion profile, but only in the analysis of the whole velocity distribution which, however, requires a large number of galaxies (of the order of several hundreds; e.g. Dejonghe 1987; Merritt 1988; Merritt \\& Gebhardt 1994). Without some information from the relative distribution of dark and galaxy components, the virial theorem places only order-of-magnitude constraints on the total mass (e.g., Merritt 1987). The usual approach is to apply the virial theorem by assuming that mass is distributed like the observed galaxies (e.g., Giuricin, Mardirossian, and Mezzetti 1982; Biviano et al. 1993). This assumption is supported by several pieces of evidence coming both from optical (e.g., Carlberg, Yee, \\& Ellingson 1997) and X-ray data (e.g., Watt et al. 1992; Durret et al. 1994; Cirimele, Nesci, \\& Trevese 1997), as well as from gravitational lensing data, which, however, suggest a smaller core radius (e.g., Narayan \\& Bartelmann 1997). The mass estimates derived from gravitational lensing phenomena are completely independent of the cluster dynamical status, but a good knowledge of cluster geometry is required in order to go from the projected mass to the cluster mass (e.g., Fort 1994). Moreover, strong lensing observations give values for the mass contained within very small cluster regions ($\\lesssim$ one hundred of kpc) and weak lensing observations are generally more reliable in providing the shape of the internal mass distribution rather than the amount of mass (e.g., Squires \\& Kaiser 1996). Up to now, few studies have dealt with wide comparisons between mass estimates obtained by different methods for the same cluster. Wu \\& Fang (1996; 1997) found that masses derived from gravitational lensing analyses are higher than those from X-ray analyses by a factor of 2, but agree with those from galaxy analyses. Indeed, mass estimates from lensing seem to agree with X-ray estimates when clusters are relaxed (e.g., Allen 1997). However, Wu \\& Fang's works concern only clusters which lie at moderate redshifts and show gravitational lensing phenomena which could be enhanced in the presence of substructures (e.g., Miralda-Escud\\'e 1993; Bartelmann, Steinmetz, \\& Weiss 1995). For nearby clusters, there is a trend to obtain larger masses from galaxy analyses than from X-ray analyses (e.g., Cowie et al. 1987; Mushotzky et al. 1995; David, Jones, \\& Forman 1995), but acceptable agreement exists in some individual cases (e.g., for the Coma cluster, Watt et al. 1992). The classical approach of the virial theorem based on measurements of discrete velocities bears re-examining owing to the large new data sets which are now becoming available for nearby clusters, i.e. the ESO Nearby Abell Clusters Survey (ENACS) by Katgert et al. (1996, 1998). Moreover, the fair level of consistency among recent estimates of velocity dispersion of member galaxies resulting from different membership assignment procedures (cf. Fadda et al. 1996, hereafter F96, and Mazure et al. 1996) makes us confident of the robustness of our approach. The aim of this work is to obtain reliable mass estimates. These mass estimates will be used in the computation of the mass function of nearby clusters (Girardi et al. 1998). The paper is organized in the following manner. We describe the data sample and our selection procedure for cluster membership assignment in \\S~2. We briefly describe the methods used to compute cluster masses by using member galaxies in \\S~3. By assuming that mass follows the galaxy distribution, we compute virial mass estimates in \\S~4, and we verify their consistency with the results of the Jeans equation in \\S~5. The strongly substructured clusters are analyzed in \\S~6. We compare our mass estimates with those derived from X-ray analyses in \\S~7. We discuss our results in \\S~8. We give a brief summary of our main results and draw our conclusions in \\S~9. Unless otherwise specified, we give errors at the 68\\% confidence level (hereafter c.l.) A Hubble constant of 100 $h$ \\ks $Mpc^{-1}$ is used throughout. ", "conclusions": "The main points of this work may be summarized as follows: i) We evaluate in a homogeneous way the optical masses of 170 nearby clusters ($z\\le 0.15$). This sample, which is the largest set of clusters up to now analyzed in the literature, includes both data from the literature and the new ENACS data (Katgert et al. 1996, 1998). ii) On the assumption that mass follows the galaxy distribution, we compute the masses of each cluster by applying the virial theorem to the member galaxies and we verify our results by using the Jeans equation. iii) Our mass estimates are smaller than previous optical estimates. This fact is due both to our better membership assignment procedure and to the application of the correction due to the presence of the surface term in the virial theorem (recently stressed by Carlberg et al. 1997a). iv) After appropriate rescaling to the X-ray radii, we compare our optical mass estimates to those derived from X-ray analyses, which we have compiled from the literature (for 66 clusters). We find a good overall agreement. v) The above agreement is expected on the basis of two common assumptions: a) that mass follows the galaxy distribution, b) that clusters are not far from a situation of dynamical equilibrium with both gas and galaxies reflecting the same underlying mass distribution. It should be pointed out that Carlberg et al. (1997a) have recently drawn similar conclusions for a sample of distant clusters (the CNOC sample). In particular, we find evidence for a galaxy distribution which is colder and less extended than the gas distribution. Several recent studies have casted doubts on cluster mass estimates and attempted to lower the cluster baryon fraction by reducing the cluster masses (e.g., Gunn \\& Thomas 1996; Wu \\& Fang 1996). We stress that our study strongly supports the reliability of present cluster mass estimates derived from X-ray analyses and/or (appropriate) optical analyses. Hence, it is even more difficult to reconcile present data with a $\\Omega_0=1$ Universe (e.g. White et al. 1993b). Our cluster masses are suitable for statistical studies. In particular, we did not reject a priori those clusters with a poor number of selected members, which usually have a small mass, in order to avoid having a final cluster sample biased towards more massive systems." }, "9804/astro-ph9804323_arXiv.txt": { "abstract": "\\he\\ \\lya\\ \\lm 304/\\ha\\ \\lm 1640 emission lines are mainly produced by recombination, and their canonical ratio of $\\sim 10$ may be a sensitive reddening indicator. We obtain the high S/N optical spectra of two quasars and combine them with the far-UV spectra that show the \\he\\ \\lm 304 emission. For HS~1700+64, the \\he\\ \\lm 1640 emission is not detected, and an upper limit to it sets the ratio greater than 20. This may not be inconsistent with the theoretical value when all observational uncertainties are taken into consideration. For Q0302-003, the ratio is very low, on the order of unity. The most plausible cause for such a low ratio is extinction in the EUV band by very fine grains of dust. Q0302-003 has a prominent narrow component of $\\rm FWHM \\sim 2000 \\ \\kms$ in its major emission lines, and it appears that reddening is associated only with the line-emitting region. We suggest that the geometry of the line-emitting region in high-z quasars resembles that in the low-luminosity active galaxies, with the presence of dust mostly in the outer part. ", "introduction": "The first UV spectroscopic observation of a quasar (3C~273, \\cite{davidsen77}) enabled a measurement of the \\lya/\\ha\\ ratio in an active galactic nucleus (AGN). The low ratio of $\\sim 1$ was in line with an independent study using the composite spectrum that was derived from various quasars (\\cite{baldwin}), but was in sharp conflict with a canonical ratio of $\\sim 10$ predicted by standard photoionization models (\\cite{deo}). This ``\\lya/\\ha\\ problem'' raised serious concerns about the validity of photoionization as the main line-emission mechanism in AGN and prompted extensive theoretical interest in the following years. Improved photoionization models (\\cite{ferland}, and references therein) invoke large column densities and moderate degrees of ionization. In a partially ionized zone, the low escape probability for \\lya\\ photons makes a high population of excited states, and collisional excitation from these levels enhances Balmer lines. Calculations using a reasonable AGN continuum lead to an enhancement of Balmer lines by a factor of $\\sim 2$, and hence may not fully explain the observed low \\lya/\\ha\\ ratio. Another explanation introduces intrinsic reddening in the line-emitting region (\\cite{dust}). The wavelength-dependent extinction reduces the intensities of observed UV lines, thus lowering the \\lya/\\ha\\ ratio. The observed Pa$\\alpha$/\\ha\\ ratio, however, appears to be too low for a straightforward full account as a reddening effect (\\cite{rick}). Significant evidence exists for dust in the narrow-line region of Seyfert galaxies (\\cite{deo}; \\cite{n93}) and for a decrease in extinction with increasing luminosity toward quasars (\\cite{cdz}; \\cite{rudy}). However, \\cite{wills} suggested that in intermediate-redshift quasars there may be significant reddening in the narrow-line region. Accurate assessments of the reddening effect depend on the use of good line pairs whose intrinsic ratios are fairly stable. Since singly ionized helium is hydrogenic, the \\he\\ I(\\lya\\ \\lm 304)/I(\\ha\\ \\lm 1640) ratio should therefore be the same as that for hydrogen. The \\he\\ emission is produced mainly by recombination, because its excitation level of 40 eV is considerably higher than the average thermal energy in the line-emitting region. The wavelength of \\he\\ \\lm 304 emission coincides with that of O$^{++}$ transition $2p^2\\ ^3P_2 - 2p\\ 3d\\ ^3P_2^0$, allowing Bowen fluorescence radiation (\\cite{eastman}; \\cite{netzer}). This radiation mechanism, however, does not appreciably affect the \\he\\ ratio itself. The \\he\\ \\lm 304 emission should be extremely sensitive to reddening. While there are no hurdles in the theoretical aspects, it has taken some 20 years to advance from measuring this ratio in hydrogen to that in helium. ", "conclusions": "The \\ratio\\ ratio is quite different in these two quasars. Indeed the line profiles in Q0302-003 are narrower, making it easier to identify the weak \\he\\ \\lm 1640 feature. The S/N level is high enough that a \\he\\ \\lm 1640 feature should be detected even with a line width of $\\sim$ 12 000 \\kms. In Fig. 1 the profile of an assumed \\he\\ \\lm 1640 feature is plotted, with an intensity 20\\% of the \\he\\ \\lm 304, which should have been detected. It appears that difference is not simply attributable to line widths. The intensity of \\he\\ \\lm 304 emission is affected by the Lyman line and continuum absorption by numerous intervening absorbers along the line of sight. This can be corrected if a high-resolution spectrum at longer wavelengths yields a list of absorption lines. Our estimate, based on the statistical result of \\cite{valley}, suggests an optical depth of 0.2 at a rest-frame wavelength of 300 \\AA\\ for a z=3.3 quasar. Therefore, this Lyman-Valley correction is not very significant. The UV and optical observations are not simultaneous, and both quasars are probably variable. A comparison of the UV spectra of HS 1700+64 obtained between 1991 and 1995 finds a significant discrepancy in flux level, and that between the optical spectra taken between 1994 and 1996 shows that the \\lya\\ equivalent width varies by a factor of 2. The \\civ\\ equivalent width of Q0302-003 has increased by $\\sim 50\\%$ as compared with the data of Sargent, Steidel \\& Boksenberg (1989). Furthermore, the photometric quality of our optical spectra is questionable. A typical light loss with a small slit during an optical spectroscopic observation is $\\sim 15\\%$. These factors add uncertainties to the \\ratio\\ ratio. The derived line ratio is also subject to the reddening formulation. If we use the formula of Burstein \\& Heiles (1978), this ratio would be even lower. We have carried out photoionization calculations (\\cite{cloudy}) with various parameters. With a broad range of the density, column density, flux and shape of the ionizing continuum, the \\ratio\\ ratio varies within a narrow range between 9 and 11. It is therefore not practical to attribute the observed low value to special conditions in the line-emitting region. Note that a part of the \\he\\ \\lm 304 emission may receive a contribution from \\oiii\\ \\lm 305 emission that is produced by Bowen fluorescence mechanism (\\cite{eastman}). If this were the case, the actual \\ratio\\ ratio would be even lower. The low \\ratio\\ ratio in Q0302-003 may signal internal reddening in the line-emitting region. Dust grains with dimensions of $\\sim 3 \\times 10^{-6}$ cm are believed to produce Galactic extinction (\\cite{ext}) which generally follows a $1/\\lambda$ law. If intrinsic extinction is produced by even smaller grains, and the wavelength dependence of the extinction law applies to wavelengths as short as 300~\\AA, then an $E_{E - V} = 2.5$, where $E$ denotes a band around 300 \\AA, is needed to explain the discrepancy between the observed and theoretical ratio of \\ratio. This would translate into $E_{B-V} = 0.5$. If this is the case, significant presence of dust may be a reality in the broad-line region of some high-redshift quasars. The quantitative formalism should be more complicated than that. While the extinction curves between 1 $\\mu$m and 1000~\\AA\\ can be approximated with a $1/\\lambda$ law, very little is known about the extinction properties below 1000~\\AA. \\cite{hawkins} and \\cite{martin} calculated the EUV extinction curve for graphite-silicate dust. Their results show {\\em decreasing} extinction from 1000~\\AA\\ to 100~\\AA. \\cite{pei} suggested that a numerical formula can be applied to other galaxies, possibly to those of higher redshifts, without assuming a Galactic dust-to-gas ratio. If such extinction is real, the same extinction affects the intensity of other UV lines as well. The intrinsic hydrogen \\lya/\\ha\\ ratio in these objects may actually be higher than the observed ones. Likely, the average ratio would be $ \\sim 6$, closer than the theoretical value of $\\sim 10$. This will, in turn, help understand the classical \\lya/\\ha\\ puzzle. If significant reddening does exist in the quasar broad-line region, the observed \\lya/\\ha\\ ratio may be corrected upward by an additional factor of $\\sim 2$. Even for these two quasars, the I(\\lm 1216)/I(\\lm 304) ratio is very different. In HS~1700+64, this ratio is about 3, while in Q0302-003 it is about 30. Significant extinction in the EUV band can explain both the abnormal line ratios. The I(\\lm 1216)/I(\\lm 304) ratio in both objects is 30 or higher, consistent with photoionization models. Therefore, the likely cause for the low ratio in Q0302-003 is extinction in the broad line region by very small grains. It may not be coincidental that narrow line widths and possibly significant reddening are present in the same object. Seyfert 2 galaxies often show a higher degree of extinction (\\cite{deo}), and the narrow-line region in Seyfert-1 galaxies generally exhibit a more significant reddening effect than the broad-line region (\\cite{luc}). The spectrum of Q0302-003 shows a significant narrow component with FWHM $\\sim 2000 $~\\kms\\ for major emission lines. Although this line width is not considered very narrow for Seyfert galaxies, it is for high-z quasars. Generally, narrow lines in high-z quasars (\\cite{sargent}) are not as common as in Seyfert galaxies. For example, \\cite{wills} found no detection of narrow line components in their radio-loud quasars of $\\rm 0.26 < z < 0.77$. They suggested a significant reddening with $E_{B-V} \\simeq 0.5$. We suggest that the geometry of the line-emitting region in high-z quasars resembles that in the low-luminosity active galaxies, with the presence of dust mostly in the outer part. Making an analogy of Seyfert galaxies and some low-redshift quasars, we suggest that Q0302-003 has a narrow-line region which contains a significant amount of dust. We suggest that the geometry of the line-emitting region in high-z quasars resembles that in the low-luminosity active galaxies, with the presence of dust mostly in the outer part (\\cite{luc}). Does reddening apply to the EUV continuum? Recent studies (\\cite{n95}; \\cite{bechtold}) found that the \\lya/\\hb\\ ratio ranges between about 1 and 40 and is approximately proportional to $f(1216)/f(4861)$, the ratio of continuum flux at adjacent points (\\cite{bechtold}). Such a correlation may suggest a possible reddening effect that applies to both the continuum and lines. In such cases, the equivalent widths of concerned lines should be fairly constant. Given the significant difference in the equivalent widths of \\he\\ lines in our quasar samples, we see no compelling reason that the continuum emission from the central source is heavily reddened. Significant reddening would also affect the intensities of infrared lines, and future studies of these lines may provide additional evidence for fine dust in the quasar environment." }, "9804/astro-ph9804115_arXiv.txt": { "abstract": "The observed map of 1.809 MeV gamma-rays from radioactive $^{26}$Al (Oberlack et al, 1996) shows clear evidence of a Galactic plane origin with an uneven distribution. We have simulated the map using a Monte Carlo technique together with simple assumptions about the spatial distributions and yields of $^{26}$Al sources (clustered core-collapse supernovae and Wolf Rayet stars; low- and high-mass AGB stars; and novae). Although observed structures (e.g., tangents to spiral arms, bars, and known star-forming regions) are not included in the model, our simulated gamma-ray distribution bears resemblance to the observed distribution. The major difference is that the model distribution has a strong smooth background along the Galactic plane from distant sources in the disk of the Galaxy. We suggest that the smooth background is to be expected, and probably has been suppressed by background subtraction in the observed map. We have also found an upper limit of $1 M_{\\sun}$ to the contribution of flux from low-yield, smoothly distributed sources (low-mass AGB stars and novae). ", "introduction": "The gamma-ray created by the decay of $^{26}$Al to $^{26}$Mg was the first discovered Galactic gamma-ray line (Mahoney et al 1982). The $^{26}$Al nucleus decays by positron emission to the first excited state of $^{26}$Mg, which subsequently decays to the ground state emitting a 1.809~MeV gamma-ray. The mean lifetime of $^{26}$Al, $\\tau = 1.05$ x $10^6$ years, makes the 1.809 MeV gamma ray line an excellent tracer for newly synthesized material released into the ISM over the last several million years. The main production mechanism of $^{26}$Al is proton capture on $^{25}$Mg. Astrophysical environments that can produce $^{26}$Al include hydrostatic H-burning in the convective cores of massive stars and the H-burning shells of intermediate mass stars, and explosive H burning in novae. The carbon and neon rich shells of massive stars are also a site for $^{26}$Al production both statically and explosively. In addition to its production, the fresh $^{26}$Al must be transported into the ISM before it decays in order to be observable. The explosive mechanisms present no problems, but the transport timescale in AGB stars is of similar order to the decay timescale causing a reduction in the amount of $^{26}$Al released into the ISM. The first map of Galactic 1.809 MeV gamma-ray emission from $^{26}$Al was published by Oberlack et al (1996) from COMPTEL data. This map has a 1$\\sigma$ angular resolution of $1.6^{\\circ}$, or $3.8^{\\circ}$ FWHM. The map was produced using a Maximum-Entropy method after background subtraction. The 1.809 MeV gamma-ray map has several important characteristics, including the concentration of emission in the Galactic plane, a strong, irregular emission region toward the inner Galaxy, and a generally uneven, or clumpy, emission distribution. Along the Galactic plane there are several disconnected emission regions, some of which have been associated with O-B associations, spiral arm tangents, and the Vela SNR. Chen et al (1996) also identify several of the regions with spiral arm tangents. A recent and thorough review of the entire topic including observation, sources, and distribution of $^{26}$Al can be found in Prantzos and Diehl (1996). These observations can best be explained with sources that are spatially concentrated and rare. If the major sources of emission had small yields and a smooth Galactic distribution, the emission would be quite uniform. This is not seen in the published results (Oberlack et al 1996), which show large gaps along the Galactic plane between emission regions. We have therefore built a Monte Carlo model for the Galactic $^{26}$Al emission containing all potential astronomical sources. We have also allowed the most massive stars to form clusters that do not dissociate in the lifetime of those stars. We have made only the simplest of assumptions about Galactic structure, an exponential disk and a bulge. We have not attempted to represent any specific observed structures in the Galaxy. All non-uniformities arise from the random nature of the simulation. This produces a map that, to the eye, has a strong resemblance to the observations, with the exception of a persistent uniform background not found in the reduced observational data. We plan to use more statistically rigorous methods to measure the strength of this similarity in future work, as well as to consider some of the effects of non-uniform structure and the enhanced resolution of the INTEGRAL observatory. ", "conclusions": "" }, "9804/astro-ph9804265_arXiv.txt": { "abstract": "The colossal power output of active galactic nuclei (AGN) is believed to be fueled by the accretion of matter onto a supermassive black hole. This central accreting region of AGN has hitherto been spatially unresolved and its structure therefore unknown. Here we propose that a previously reported `deep minimum' in the X-ray intensity of the AGN MCG$-$6$-$30$-$15, was due to a unique X-ray occultation event and that it probes structure of the central engine on scales $< 10^{14} \\ \\rm cm$, or $1.4\\times 10^{-7}$ arcseconds. This resolution is more than a factor of $\\sim 3\\times 10^{6}$ greater than is possible with current X-ray optics. The data are consistent with a bright central source surrounded by a less intense ring, which we identify with the inner edge of an accretion disk. These may be the first direct measurements of the spatial structure and geometry of the accreting black-hole system in an active galaxy. We estimate a mass lower limit for sub-Eddington accretion of $3.1\\times 10^{5} M_{\\odot}$. If the ring of X-ray emission is identified with the inner edge of an accretion disk, we get mass upper limits of $1.9\\times10^{8}$ and $9.1\\times10^{8} M_{\\odot}$ for a non-rotating and maximally rotating black hole respectively. We point out that our occultation interpretation is controversial in the sense that X-ray variability in AGNs is normally attributed to intrinsic physical changes in the X-ray emission region, such as disk or coronal instabilities. ", "introduction": "\\label{intro} The accretion of matter onto a supermassive black hole as a mechanism for fueling the output of active galactic nuclei (AGN) is a paradigm strongly supported by recent spectroscopic observations of the iron K$\\alpha$ X-ray emission line (Tanaka \\etal 1995; Yaqoob \\etal 1995; Nandra \\etal 1997a and references therein). The extreme Doppler and gravitational energy shifts of the line photons, together with the shape of the line are consistent with an origin in a disk rotating about a black hole (Fabian \\etal 1989; Laor 1991). Strong gravitational redshifts, in which photon energies are changed by more than 10\\%, occur only when matter approaches closer than $\\sim 20$ gravitational radii ($= 20r_{g}; r_{g} \\equiv GM/c^{2}$) from a compact object. However, it is not possible to directly map the physical structure of the system since the highest spatial resolution of X-ray optics technology is a factor $\\sim 10^{6}$ too poor for even the closest AGN. Optical and radio observations provide greater resolution but the bulk of the emission at these wavelengths is not generated close enough to the central engine. So far, the highest spatial resolution observations, at radio wavelengths, have revealed a Keplerian disk in the AGN NGC 4258 down to only $\\sim 60,000 r_{g}$ (Miyoshi \\etal 1995; Maoz 1995). This still falls short by a factor $\\sim 3000$ of mapping the black-hole region. The AGN MCG$-$6$-$30$-$15 ($z= 0.008$) was observed by the X-ray astronomy satellite {\\it ASCA} ({\\it Advanced Satellite for Astrophysics and Cosmology}; Tanaka, Inoue, and Holt 1994) for $\\sim 4.2$ days on 23 July 1994. Results from this observation have already appeared in the literature, including the 0.5--10 keV lightcurve (Iwasawa \\etal 1996, hereafter I96; Reynolds 1997; Yaqoob \\etal 1997; see Figure 1) and a broad, asymmetric, variable iron K line with a strong red wing, consistent with a disk inclined at $\\sim 30^{\\circ}$ rotating about a black hole (Tanaka \\etal 1995; I96). The X-ray luminosity exhibits erratic variability on all timescales down to $<50$ s (Matsuoka \\etal 1990; Green, McHardy, and Lehto 1995; Reynolds \\etal 1995; Nandra \\etal 1997b). Causality arguments alone cannot put constraints on the size of the X-ray emission region since the high-frequency, lower amplitude variability may occur at localized regions of the source. Figure 1 shows an extended intensity dip at the end of the observation, from $\\sim 3.3\\times10^{5}$s to $\\sim 3.6\\times10^{5}$s. This feature has previously been dubbed as the 'deep minimum', or DM (I96). A closer inspection (Figure 2a) reveals a remarkable (albeit approximate) symmetry about the minimum luminosity. We propose that the dip was caused by an occultation of the X-ray source by optically-thick matter. This interpretation is controversial and 'non-standard', as X-ray variability in AGNs is normally attributed to intrinsic properties of the X-ray emission region, such as disk or coronal instabilities. However, so little is know about structure of the central engine in AGNs that the occultation scenario should be explored. The obscurer must be optically thick because the dip continuum spectrum only shows evidence for {\\it weak} absorption, nowhere near enough to explain the observed intensity variation over the whole \\asca bandpass (see Weaver \\& Yaqoob 1998, hereafter WY98). The luminosity at the absolute minimum of the dip is $\\sim 0.4$ of the pre-dip value and must represent persistent emission which has much smaller surface brightness than the primary source. Hereafter we will refer only to the primary X-ray emission, unless explicitly referring to the persistent emission. The proposed obscurer very likely hides the most compact and variable part of the X-ray source, since the usual rapid variability outside the dip is absent during the obscuration. Of course, it is possible that the dip is due to intrinsic variation of the X-ray source. However, the origin of X-ray variability in AGN is not understood. Models which come close to successfully reproducing the observable quantities obtained from AGN lightcurves are of the shot-noise or `rotating hot-spot' variety (e.g. Green \\etal 1993; Bao and Abramowicz 1996 and references therein). However, the parameters of such models must be highly tuned in order to reproduce AGN power spectra. On the other hand, we show in this paper that a very simple-minded model can account for the temporal profile of the intensity dip in MCG $-$6$-$30$-$15 and briefly discuss the implications for AGN X-ray variability in general. ", "conclusions": "If accreting matter is exposed to the same UV/X-ray luminosity that we observe ($L \\sim 4 \\times 10^{43} \\ \\rm ergs \\ s^{-1}$) then for gravitational infall to overcome outward radiation pressure requires the mass of the central black hole, $M_{\\rm BH}$, to exceed $3 \\times 10^{5} M_{\\odot}$. Thus, if $r_{1}$ is the radius of the inner edge of the accretion disk, identified with the last stable orbit of matter, and $\\Delta t$ is the time taken for the obscurer to traverse $r_{1}$ (i.e. $t_{3}-0.5[t_{2}+t_{1}$]), then $\\kappa r_{g}=r_{1}$ where $\\kappa=6$ or 1.24 for Schwarzschild or maximally rotating Kerr metrics respectively. But, $ r_{1} 3 \\times 10^{5} M_{\\odot}$ implies $d<1.8\\times 10^{16}\\ \\rm cm$ and $d<4.2\\times10^{17}\\ \\rm cm$ for Schwarzschild and Kerr metrics respectively. The origin of the optically-thick blobs is unspecified but Guilbert and Rees (1988) presented some simple arguments for the existence of dense ($n > 10^{15} \\rm \\ cm^{-3}$), optically thick matter residing at the heart of accreting sources. The only independent estimate of the size of the `blobs' is that they should be much thicker than $10^{9}/n_{15} \\ \\rm cm$ ($n_{15}$ in units of $10^{15} \\rm \\ cm^{-3}$). The blobs must be optically thick even near their physical boundaries (i.e they must have fairly sharp edges), otherwise the dip profile would not be so well defined. Also, the blobs must be fairly stable, esepcially if they are created in the central region itself and 'propelled' up to high altitudes. The nature of the bright central X-ray source is intriguing. An inclined jet is unlikely, since even at $30^{\\circ}$ the inflexions in the dip profile ($t_{2}$--$t_{3}$ and $t_{6}$-$t_{7}$) would have different durations. If the central X-ray source extracts its energy directly from the black hole then the metric is likely to be Kerr since energy cannot be extracted from a non-rotating black hole (Blandford and Znajek 1977; see also Ghosh and Abramowicz 1997). Occultations such as the one described here may occur frequently in AGN, but the relative sizes of the obscurer and source must be just right in order to observe such a clear event. Indeed the usual rapid variability or flicker, may be partly caused by the transit of optically thick bodies smaller than the source. We tested this hypothesis, again using a simple-minded model. Representing the bright central source as a circular disk with uniform emissivity (ignoring the ring due to its weaker emission), optically-thick blobs with ranges in radii (relative to the source) and velocities taken from Gaussian distributions were passed over the source. An additional parameter is required to specify the 'blob birth-rate' (i.e. the rate at which new blob trajectories are started). Such a model was used to produce predicted lightcurves for different model parameters. The power spectrum of each lightcurve was computed using the method of Papadakis and Lawerence (1993), omitting Poisson noise. In the range $\\sim 10^{-5}$ Hz to $\\sim 10^{-2}$ Hz, no preferred or 'universal' power-law spectral slope was found. It is possible to produce power-law spectra with slopes similar to those typically measured ($\\sim -1$ to $-2$) but for most parameter values, the slopes are too steep. Thus, fine-tuning would be necessary to explain the `universal' power-law slopes found in the handful of AGN in which it can be measured (e.g. Lawerence and Papadakis 1993; Green \\etal 1993). This is essentially because if the product of blob crossing-time and birth-rate is too large or too small, there will be no variability. The direct simulated lightcurves assume, of course that the {\\it intrinsic} source intensity is constant, which almost certainly is not the case. Thus our model does not explain AGN variability in general but its effects are potentially important to consider in any model of AGN variability. We thank the \\asca TEAM and mission operations at ISAS, Japan, for their efforts and hard work; Kim Weaver for her work on the dip spectrum, and Karen Leighly, Paul Nandra for some useful discussions. We also thank the anonymous referee. This research made use of archival data at the HEASARC, Laboratory for High Energy Astrophysics, NASA/Goddard Space Flight Center." }, "9804/astro-ph9804053_arXiv.txt": { "abstract": "We investigate different types of neutrino hot dark matter with respect to structure formation and anisotropies in the cosmic microwave background radiation (CMBR). The possibility of neutrino hot dark matter produced through the decay of a heavier neutrino by the process $\\nu_H \\to \\nu_L + \\phi$, where $\\phi$ is a scalar particle, is discussed in detail. This type of dark matter can possibly be distinguished observationally from the standard neutrino dark matter by using new CMBR data from the upcoming satellite missions MAP and PLANCK. ", "introduction": " ", "conclusions": "" }, "9804/astro-ph9804029_arXiv.txt": { "abstract": "The importance of the interstellar magnetic field is studied in relation to the evolutions of superbubbles with a three-dimensional (3D) numerical magnetohydrodynamical (MHD) simulation. A superbubble is a large supernova remnant driven by sequential supernova explosions in an OB association. Its evolution is affected by the density stratification in the galactic disk. After the size reaches 2--3 times the density scale-height, the superbubble expands preferentially in the $z$-direction. Finally it can punch out the gas disk (blow-out). On the other hand, the magnetic field running parallel to the galactic disk has an effect to prevent from expanding in the direction perpendicular to the field. The density stratification and the magnetic fields have completely opposite effects on the evolution of the superbubble. We present results of 3D MHD simulation in which both effects are included. As a result, it is concluded that when the magnetic field has a much larger scale-height than the density, even for a model that the bubble would blow out from the disk if the magnetic field were absent, the magnetic field with the strength of 5 $\\mu$G can confine the bubble in $|z| \\alt 300$ pc for $\\simeq$20 Myr (confinement). In a model that the field strength decreases in the halo in proportion to $B \\propto \\rho^{1/2}$, the superbubble eventually blows out like a model of $B=0$ even if the magnetic field in the mid-plane is as strong as $B=5\\mu$G. ", "introduction": "A superbubble is a complex consisting of an OB association, surrounding X-ray emitting hot gas, and a corresponding HI hole/shell. Three examples are well-known in our Galaxy: Cygnus (Cash et al. 1980), Orion-Eridanus (Cowie, Songaila \\& York 1979; Reynolds \\& Ogden 1979), and Gum nebula (Reynolds 1976). Superbubbles are found as HI shells and holes in external galaxies, such as LMC (Meaburn 1980; Dopita, Mathewson \\& Ford 1985), M31 (Brinks \\& Bajaja 1986), M33 (Deul \\& den Hartog 1990), M101 (Kamphuis, Sancisi \\& van der Hulst 1991), and so on. Since the sizes of these objects are in the range of 100 pc -- 1kpc, this can not be explained by a single supernova explosion [the size of an ordinary supernova remnant (SNR) is $\\alt 50$pc]. The amount of energy required for such a superbubble reaches $5\\times 10^{51} {\\rm erg} - 10^{54} {\\rm erg}$ (Tenorio-Tagle \\& Bodenheimer 1988). There are two models proposed for formation of superbubbles: (1) a large SNR driven by sequential supernova explosions in an OB association and (2) a complex formed by a collision of a high-velocity cloud and the galactic disk (Tenorio-Tagle 1991). Here, we confine ourselves to the first model and discuss the evolutions. For review papers of this field, see Tenorio-Tagle \\& Bohdenheimer (1988), Spitzer (1990), Tomisaka (1991), Bisnovatyi-Kogan \\& Silich (1995). Here, we summarize the evolution very briefly. After the size of bubble exceeds the density scale-height, the bubble becomes elongated in the direction perpendicular to the galactic disk (Tomisaka \\& Ikeuchi 1986; Tenorio-Tagle, Bodenheimer \\& R\\'{o}\\.{z}yczka, 1987; MacLow \\& McCray 1987). It is shown that when the mechanical luminosity released by sequential supernova explosions is high, the expansion of the bubble to the halo is accelerated and the hot gas contained in it flows into the galactic halo (galactic fountain), finally. However, the magnetic fields running parallel to the galactic disk prevents the gas from flowing in the vertical direction (perpendicular to the fields). Tomisaka (1990) studied the adiabatic evolution of a superbubble in uniform magnetic fields and showed that the bubble driven by a mechanical luminosity of $L_{\\rm SN}\\simeq 3\\times 10^{37} {\\rm erg\\,s^{-1}}$ is confined in the galactic disk by the effect of the magnetic field provided its strength is as large as $B_0=5\\mu$G. This shows the superbubble is confined in the galactic disk, if (1) the magnetic fields have a large scale-height, $H_B$, and (2) $L_{\\rm SN}\\simeq 3\\times 10^{37}{\\rm erg s^{-1}}$ and $B_0\\agt 5\\mu$G. However, this result may be affected by the assumption of the adiabatic gas. The interstellar magnetic fields seems to play an important role in the evolution of a superbubble (see also Mineshige, Shibata, \\& Shapiro 1993, Ferriere, MacLow, \\& Zweibel 1991). In the present paper, a full magnetohydrodynamical calculation has been done including the radiative cooling. Further, the effect of distributions of magnetic field strength is studied. Plan of the present paper is as follows: in section 2 is given an analytical estimate to a threshold mechanical luminosity under which the superbubble is confined in the galactic disk. From this estimation, we choose the values for the parameters $L_{\\rm SN}$, $B_0$, and $H_B$. Numerical method and other assumptions are also presented in $\\S$2. Section 3 is for the numerical result, in which the effect of the magnetic fields is shown. In section 4, we will discuss the observability of the superbubble in external galaxies. Using the evolution obtained, we will show that a large fraction of the interstellar space is occupied with superbubbles. ", "conclusions": "\\subsection{Blow-out or confinement?} As shown in the preceding section, although in Model B the mechanical luminosity is much larger than the critical luminosity of equation(\\ref{eqn:Lcr0}), the vertical expansion is much decelerated by the effect of magnetic fields. This shows that $L_{\\rm SN} < L_{\\rm crit}$ may be only a sufficient condition for confinement of the superbubble. The expansion of a spherical shell driven by a pressure in the hot cavity $p$ is formulated as follows: \\begin{equation} \\frac{dMv}{dt}=4\\pi\\, R^2 (p-p_{\\rm out}), \\end{equation} \\begin{equation} \\frac{dM}{dt}=4\\pi\\, R^2\\, v\\, \\rho_{\\rm out} \\end{equation} \\begin{equation} \\frac{dR}{dt}=v, \\end{equation} \\begin{equation} \\frac{dE}{dt}=L_{\\rm SN}-4\\pi R^2\\, p\\, v, \\end{equation} where, $M$, $R$, $v$, $p_{\\rm out}$, $\\rho_{\\rm out}$ and $L_{\\rm SN}$ represent, respectively, the mass of the shell, the radius and velocity of the shell, the interstellar pressure and its density and the energy release rate from an OB association. These equations are, respectively, the equation of motion, the mass conservation, the relation of a size $R$ to a velocity $v$, and the first law of thermal physics. If we assume $p=(2/3)\\times E /(4\\pi R^3/3)$, these four equations can be solved numerically. Figure 9 shows a resultant expansion law of a spherical superbubble with $L_{\\rm SN}=3\\times 10^{37}{\\rm erg~s^{-1}}$ in a {\\em uniform interstellar medium} of $n_0=0.3{\\rm cm^{-3}}$. Each curve corresponds to different external pressures as $p_{\\rm out}=1.7\\times 10^{-12}{\\rm erg~cm^{-3}}$ (solid line), $p_{\\rm out}=1\\times 10^{-12}{\\rm erg~cm^{-3}}$ (dotted line), and $p_{\\rm out}=0$ (dashed line). Weaver et al.'s (1977) solution, equation (1), agrees with the curve of $p_{\\rm out}=0$. This figure shows that the interstellar pressure plays an important role especially in the late phase of the evolution. This is understood as follows: in the late phase of the superbubble the difference between the internal pressure and the outer one is small and this small difference drives the shell further. Therefore, the critical luminosity would be underestimated if we use a solution without taking the outer pressure into account. The shell of a superbubble continues to expand as long as the energy ejection continues, while a supernova remnant stops its expansion after $p_{\\rm out}=p$. Thus, exactly speaking, the superbubble is never confined as long as the OB association is alive. However, a slow expansion driven by a small pressure difference as $p-p_{\\rm out}$ is considered as a signature of the confinement. In this figure, we also plot an equivalent radius, which is defined using the volume occupied with a hot matter ($V_{\\rm hot}$) as \\begin{equation} R_{\\rm equiv}\\equiv \\left(\\frac{3V_{\\rm hot}}{4\\pi}\\right)^{1/3}, \\label{Requiv} \\end{equation} for Models A and C. The equivalent radius for Model A in which the bubble is almost confined in the galactic gaseous disk shows a similar expansion law to that obtained by a thin-shell model (a solid curve). In contrast, that of Model C indicates a completely different expansion law such that after $t \\agt 10$ Myr the equivalent radius increases rapidly and in $t\\simeq 35$Myr $R_{\\rm equiv}$ surpasses the model with $p_{\\rm out}=0$. These differences seem to come from the distribution of magnetic field strength. In Model A the total pressure (thermal plus magnetic one) is almost constant as the bubble expands, because the magnetic pressure is dominant over the thermal one and magnetic fields are uniform. While, in Model C the total pressure drops according to the density distribution $\\rho(z)$. This figure indicates that when the equivalent radius is well fitted by this thin shell model the bubble is nearly confined in the disk even if the gas disk has a finite scale-height. In contrast, if hot gas is ejected from the galactic disk, the equivalent radius shows a more rapid expansion than that derived by this thin-shell model. Values of equivalent radii at the epochs when numerical runs end are shown in table 1. \\subsection{Observability} A shear motion in the galactic rotation and rotation itself may play a role in the evolution of a superbubble (Tenorio-Tagle \\& Palou\\v{s} 1987; Palou\\v{s} et al. 1990; Silich 1993). Galactic shear seems to deform the shape and the Coriolis force makes the shell rotate. The characteristic time-scales of the rotation, $\\tau_R$, and the shear, $\\tau_S$ are estimated respectively as \\begin{equation} \\tau_{\\rm R} \\sim 1/\\Omega_0 \\sim 40{\\rm Myr} (\\Omega_0/26{\\rm km~s^{-1}~kpc^{-1}})^{-1} \\end{equation} and \\begin{equation} \\tau_{\\rm S}\\sim (l d\\Omega/dR)^{-1} \\sim 320 {\\rm Myr} (l/1{\\rm kpc})^{-1}(\\Omega_0/ 26{\\rm km~s^{-1}~kpc^{-1}})^{-1} (R_0/8.5{\\rm kpc}), \\end{equation} where $\\Omega_0$, $R_0$, and $l$ are the angular speed of galactic rotation, distance from the galactic center, and a typical size of a superbubble, respectively. Since active SN-explosion phase continues for $\\sim 50$ Myr for an OB association (McCray \\& Kafatos 1987), in the late phase of $\\tau_{\\rm R}\\la t \\la 50$Myr, the effect of the Coriolis force seems to appear as a deformation force of the shell. The $\\alpha\\omega$-dynamo mechanism driven by superbubble was studied recently by Ferri\\`{e}re (1992). The galactic rotation has a little effect on the evolution of a superbubble. Thus, if shells or holes observed in external galaxies are elongated after their inclinations are corrected, their direction seems to indicate that of the magnetic fields. There have been listed 141 HI holes in M31 by Brinks \\& Bajaja (1986). These holes are observed, more or less, as elliptical. Particularly, the holes found near the major axis of M31 are important to determine the physical shape of the holes. Many of these HI holes have their major axes perpendicular to the galaxy's major axis. Since this is not explained by projection due to the inclination of M31 ($i=77\\deg$), these seem to have physically a shape elongated along the azimuthal direction of the galaxy. This is not inconsistent with observations indicating that a global pattern of the magnetic field is ring-like in M31, which is measured by radio linear polarization observations (for a review, see Sofue, Fujimoto \\& Wielebinski 1986), in other words, magnetic field lines run in the azimuth direction in M31. \\subsection{Porosity} If hot gas contained in superbubbles occupies a large volume of the galactic disk, a picture of the interstellar medium should be changed (McKee \\& Ostriker 1977). The fraction of areas covered by superbubbles younger than $\\tau_{\\rm active}$ is estimated with a quantity called as two-dimensional porosity which is defined as \\begin{equation} Q(t < \\tau_{\\rm active}) \\equiv r_{\\rm OB}\\int_0^{\\tau_{\\rm active}} S(t)dt, \\end{equation} where $r_{\\rm OB}$ is the formation rate of OB associations per unit area and $S(t)$ represents the area which covered by a hot cavity on the mid-plane of the disk $z=0$. This is identical with a two-dimensional porosity parameter calculated by Heiles (1990). He estimated galaxy-wide average of two-dimensional porosity $Q_{\\rm 2D}\\simeq 0.30$. $\\tau_{\\rm active}$ should be chosen equal to the oldest age of a superbubble which contains a hot gas inside. If we assume $\\tau_{\\rm active}=20$ Myr and integrate $S(t)$ for Models A and C ($S\\equiv \\pi x_c y_c$), these two models give respectively $2.08\\times 10^6\\, {\\rm pc^2~Myr}$ and $1.57\\times 10^6\\,{\\rm pc^2~Myr}$. We adopt the estimation of $r_{\\rm OB}$ from a galactic type II supernova rate of $r_{\\rm II}\\sim 0.01{\\rm yr^{-1}}$, that is, we assume that all type II SNe occur in OB associations, number of type II SNe in an OB association is constant irrespective of richness of association as $N_{\\rm SN}\\sim 100$ and OB associations are uniformly distributed in the galactic disk with radius $R_{\\rm gal}\\simeq 10$ kpc. This gives an estimation of OB association formation rate as \\begin{equation} r_{\\rm OB} = \\frac{r_{\\rm II}}{N_{\\rm SN}\\,\\pi\\, R_{\\rm gal}^2}, \\end{equation} \\begin{equation} r_{\\rm OB} \\simeq 3.2 \\times 10^{-7}{\\rm pc^{-2}\\,Myr^{-1}} \\left( \\frac{r_{\\rm II}}{0.01{\\rm yr^{-1}}}\\right) \\left( \\frac{N_{\\rm SN}}{100} \\right)^{-1} \\left( \\frac{R_{\\rm gal}}{10 {\\rm kpc}}\\right)^{-2}. \\end{equation} This indicates the two-dimensional porosity to be equal to $Q(t < 20 {\\rm Myr})\\simeq 0.5-0.6$. Thus, rather large fraction of the galactic disk, $1-\\exp{(-Q)}\\sim 40\\%-45\\%$, is covered by young ($t < 20$ Myr) superbubbles." }, "9804/astro-ph9804047_arXiv.txt": { "abstract": "We consider magnetic field evolution of neutron stars during polar-cap accretion. The size of the polar cap increases as the field decays, and is set by the last open field line before the accretion disk. Below the polar cap we find the temperature to be so high that electron-phonon scattering dominates the conductivity. Outside the polar cap region, the temperature is such the conductivity is dominated by temperature independent impurity scattering which can be a few orders of magnitude larger than the electron-phonon conductivity. The time-scale for field decay is therefore initially given by impurity scattering dominated conductivity. When the field strength has been reduced to $\\sim 10^8 ~{\\rm gauss}$ the accretion is spherical and the time scale for field decay is given by the smaller electron-phonon scattering conductivity. The field strength is now reduced rapidly compared to before and this could be a reason for there being no pulsars known with field strengths below $10^8~{\\rm gauss}$. We also investigate the evolution of multipoles at the neutron star surface. We find that contribution from higher-order multipoles are at most 30 \\% to that of the dipole mode. ", "introduction": "Since the discovery of pulsars there has been much discussion of observational evidence for decay of magnetic fields in neutron stars, as well as much theoretical work. As the reviews by Lamb (1991), Chanmugam (1992), and Phinney \\& Kulkarni (1994) indicate, there is at present no consensus on the question of whether or not magnetic fields in isolated neutron stars can decay significantly. The general view has been that the electrical conductivity of matter in the cores of neutron stars is so high that the characteristic decay time for fields generated by electrical currents in the core is greater than the age of the Universe. Recently Pethick \\& Sahrling (1995) showed that even if the conductivity in the core was small the shortest possible decay time is some two orders of magnitude longer than the decay time for configurations where the magnetic field is confined to the crust. An incorporation of general relativistic effects, Sengupta (1997), further reduces the decay rate. Still, millisecond pulsars have typical field strengths in the range $10^8~ -~ 10^{10}~{\\rm gauss}$ compared to isolated radio pulsars which have typically field strengths around $10^{12}~{\\rm gauss}$ and this indicates that during the spin-up phase the accretion process is reducing the field strength somehow. Millisecond pulsars are generally found in binary systems where the companion star is a white dwarf with a mass less than a solar mass, $M_{\\sun}$. This system is called a Low-Mass-Binary Pulsar (LMBP) referring to the mass of the companion star. The progenitor to this system is thought to be the Low-Mass-Xray Binaries (LMXB) where a neutron star is accreting matter from a companion having a mass less than about 2 $M_{\\sun}$. The accreting matter is spinning up the neutron star to millisecond periods. For details concerning this process see reviews by Phinney \\& Kulkarni (1994) and Bhattacharya \\& van den Heuvel (1991) among others. The evolution of the binary system after the neutron star is formed either by a supernovae or tidal capture, is assumed to occur on at least two time scales when the companion star is in radiative equilibrium. At first the companion star is evolving on a nuclear time scale slowly filling its Roche lobe. When it has been filled up the matter overflows and falls onto the companion neutron star on a thermal time scale $\\tau_{th}=G M^2/R L = 5\\times10^7~ (M_{\\sun}/M)^2 ~{\\rm yrs}$, see Bhattacharya \\& van den Heuvel (1991) for details. If LMXB's are the only progenitors to LMBP's the lifetime of the LMXB, or the accretion phase of the binary system, must be of order $10^7~{\\rm yrs}$. However, by using the amount of mass needed to be accreted to spin up the neutron star to a spin period $P_i$ one finds a time scale that ranges from $10^8-10^{10}\\times (P_i/2~{\\rm ms} )^{-4/3}~{\\rm yrs}$. This discrepancy suggests that there might other progenitors to the LMBP's. For details see the review by Phinney and Kulkarni (1994). We will in this paper assume the accretion phase of progenitors to millisecond pulsars to last between roughly $10^7-10^9~{\\rm yrs}$. Matter accreting onto a neutron star is expected to be disrupted by the magnetic field at some radius, $r_A$, sometimes called the Alfv\\'en radius. Beyond this bare description there is no generally accepted view on how or where the matter attaches to the field lines and flows to the neutron star's surface, or on the interaction of the field with the matter outside $r_A$, despite a large number of papers on the subject, see King (1995). In the case of disk accretion there are two main approaches to the problem. In one (Ghosh \\& Lamb 1978, 1979a,b; Kaburaki 1986 and Wang 1987) the field is assumed to thread a large fraction of the disk because of Kelvin-Helmholtz instabilities. The other approach (Aly 1980; Anzer \\& B\\\"orner 1980, 1983; Scharlemann 1978) assumes the disk is a perfect conductor, completely excluding the field. In both approaches the matter is often assumed to leave the disk in i a narrow transition zone at the inner edge (near $r_A$), thereafter flowing along field lines to the neutron star. Most studies concerned with the magnetic field evolution {\\it inside} the neutron star have assumed matter to accrete onto the surface spherically, e.g. Fujimoto et al. (1984), Miralda-Escud\\'e et al. (1990), hereafter Mir90, Urpin \\& Geppert (1995), and Konar \\& Bhattacharya (1997). In this paper I consider non-spherical accretion where matter is assumed to flow onto the polar caps in a column. The cap is consequently heated up and the conductivity in the crust below the polar cap we estimate to be much smaller than the conductivity outside the accretion column. The magnetic evolution in this scenario can be divided into two stages where in stage I the global decay rate is controlled by the conductivity outside the accretion column, $\\tau_B\\sim 10^{8}-10^{10}~{\\rm yrs}$. When the field has reached a value of about $10^8 {\\rm gauss}$ the accretion is spherical and the evolution enters the second stage. Here, the whole crust is being heated up and the conductivity is dominated by electron-phonon scattering. In this stage the magnetic field decay time is a few orders of magnitude shorter than in stage I and compared to the earlier evolution the field is dissipating rapidly. I argue that this could account for the fact that no binary pulsars have been found with a magnetic field less than $10^8~{\\rm gauss}$. A similar effect where the accreted flow is pushing the field lines has already been noticed by Romani (1993). The paper is organized as follows: section 2 contains a brief description of the model and the basic equations. In section 3 we estimate the length scale for temperature change at the accretion column-normal crust boundary to be less than a crust thickness. The time scale for the temperature to reach a stationary state is also shown to be much smaller than the magnetic field decay time scale. Therefore, we do not solve the energy equation explicitly but assume the conductivity as a function of angle to be close to a top-hat function at all times. Section 4 presents the numerical results for some initial depths of the magnetic field. ", "conclusions": "We have examined the influence of asymmetric accretion on the magnetic field evolution of a neutron star. The temperature structure in the crust resulting from the accretion was roughly estimated and found to vary more rapidly than the crust thickness around the edge of the accretion cap. The time scale for reaching a steady state was shown to be shorter than the time scale for magnetic field decay. Therefore, instead of doing a full blown calculation of heat conduction coupled with the magnetic field evolution we used a simple smoothed top-hat function for the temperature structure and consequently for the conductivity. The global field decay time is roughly the one given by the largest conductivity in the crust, which occurs outside the accretion cap, $\\tau_{B,i} = 5.5 \\times 10^{17}~(\\delta R_5)^2~(\\rho_{14} x/0.1)^{1/3}Z/(60~ (Q/0.01))~{\\rm s}$. However, when the field strength is down to roughly $10^8$ gauss the accretion is spherically symmetric and we have $\\tau_{B,ph} = 5.5 \\times 10^{15}~ (\\delta R_5)^2~\\rho_{14}^{7/6} ~(x/0.1)^{5/3}T_8^{-2}~ {\\rm s} \\ll \\tau_{B,i}$ and so the decay rate is much faster than initially which could provide a clue why no binary pulsars are known with field strengths less than $10^8$ gauss. Romani (1993) discussed a similar effect and found a threshold close to ours. We also found the asymmetric accretion resulted in some higher order surface magnetic multipoles and the size of these were shown to be at most 30 \\% of the surface dipole field. {\\bf Acknowledgements :} I would like to thank Dong Lai, Lars Bildsten and Edward Brown for useful discussions. Dong Lai is especially thanked for a careful reading of the manuscript. This work was also supported in part by the U. S. National Science Foundation under grants NSF AST93-15133 and AST94-14232, by NASA under grant NAGW-1583, and by the Swedish Natural Science Research Council. \\vfill\\eject" }, "9804/astro-ph9804271_arXiv.txt": { "abstract": "The lack of bright host galaxies in several recently examined gamma--ray burst (GRB) error boxes suggests that the redshifts of cosmological GRBs may be significantly higher than previously hypothesized. On the other hand, the non--detection of multiple images in the BATSE 4B catalog implies an upper limit to the average redshift $\\langle z\\rangle$ of GRBs. Here, we calculate an upper limit to $\\langle z\\rangle$, independent of the physical model for GRBs, using a new statistical lensing method that removes distance ambiguities, and thus permits accurate computation of the lensing rate at high $z$. The upper limit on $\\langle z\\rangle$ depends directly on the cosmological parameters $\\Omega$ and $\\Lambda$. If there are no multiple images among the brightest 80\\% of the first 1802 bursts in the BATSE 4B catalog, then, at the 95\\% confidence level, $\\langle z\\rangle<$ 2.2, 2.8, 4.3, or 5.3 for ($\\Omega$, $\\Lambda$) values of (0.3, 0.7), (0.5, 0.5), (0.5, 0.0), or (1.0, 0.0), respectively. The 68\\% upper limit to the average redshift is comparable to or less than the median redshift of GRBs in scenarios in which the GRB rate is proportional to the rate of star formation, for any cosmology. The uncertainty in the lensing rate---arising from uncertainties in the cosmological parameters and in the number density and average velocity dispersion of galaxies---will be reduced significantly in the next few years by a new generation of experiments and surveys. Moreover, the continued increase in the number of GRBs observed by BATSE will greatly constrain their redshift distribution. ", "introduction": "Three decades after their discovery (\\cite{KSO73}), the physical origin of gamma--ray bursts (GRBs) remains unresolved. Recent developments, however---including the isotropy of GRBs seen by BATSE (\\cite{B96}; \\cite{THBM96}) and the detection of redshift $z$=0.835 absorption and emission lines from a possible optical counterpart to GRB 970508 (\\cite{M97}; \\cite{R98})---strongly suggest a cosmological origin for GRBs. Initially, no--evolution fits to the log N -- log P (peak flux) distribution of BATSE GRBs suggested a typical redshift of $z\\sim 1$ for dim bursts (\\cite{F93}; \\cite{W93}), with the break in the $-3/2$~slope of the flux distribution at $P \\sim 10$ ph~cm$^{-2}$~s$^{-1}$ being interpreted as a cosmological deviation from Euclidean space at $z \\sim 1$. In addition, a number of researchers reported a factor of $\\sim 2$ ``time--stretching'' of dim bursts relative to bright ones (e.g., \\cite{Nor94}, 1995), which was thought to imply a redshift of order unity for the dim bursts. Such a redshift would imply an extremely low rate of gravitational lensing and multiple imaging of sources detected with BATSE, perhaps as low as one multiple imaging event per 200 years (\\cite{GN94}). New lines of evidence now indicate that the typical redshift of cosmological GRBs may substantially exceed unity. Particularly suggestive evidence for high GRB redshifts comes from deep HST searches of the error boxes of several of the brightest GRBs detected by BATSE and the Interplanetary network (IPN): In five cases, no obvious host galaxies for the bursts were found, down to a limiting magnitude between 3.5 and 5.5 mag fainter than what would be expected if GRBs reside in $L_\\star$ galaxies and dim bursts are at $z\\sim 1$ (\\cite{Sch97}). This ``no--host problem'' has been analyzed further by Band \\& Hartmann (1998), and implies that if GRBs reside in normal galaxies, even the {\\em brightest} of them are at redshifts close to unity, with the faintest being at much higher redshift still. In addition, Fenimore \\& Bloom (1995) showed that the intrinsic anticorrelation between photon energy and burst duration implies that if the observed time--stretching is caused by time dilation, then the redshift of the dimmest bursts may be as large as $z\\sim 6$ (note, however, that some or all of the time--stretching may be intrinsic to the bursts; cf. \\cite{SPS97}). These new lines of evidence have led researchers to explore scenarios in which the bursts are at much higher redshifts. One currently popular scenario, motivated in part by the assumption that cosmological GRBs involve remnants of massive stars, is that the GRB rate is proportional to the rate of (massive) star formation (\\cite{Tot97}; \\cite{WBBN98}). The star formation rate is thought to vary strongly with redshift (\\cite{Ma96}), peaking at $z \\sim 2$. If this is the case, the dimmest bursts may be at redshift $z\\sim 6$, making them the most distant objects ever detected (\\cite{WBBN98}). Other authors have noted that, if the comoving number density of GRB sources is allowed to vary with redshift, then even if the peak luminosity is fixed, the log N--log P distribution and other properties of the BATSE bursts can be fit by models with a maximum redshift up to $z\\sim$10--200 (\\cite{RHL95}). In these high-$z$ scenarios, the expected incidence of gravitational lensing and multiple imaging of GRB sources is much higher than it is in lower-$z$ scenarios, since in general the lensing rate increases markedly with the source redshift. The short duration of GRBs (typically tens of seconds) compared to the difference in light travel time between different ray paths (typically months, if the lens is of galactic mass; cf. \\cite{M92}) means that a multiply-imaged GRB appears as two or more separate events with identical time histories and intensities that differ only by a scale factor. The image separations, of order arcseconds, are tiny compared to BATSE location errors; thus, these GRBs would appear to come from the same location. Two or more such events must be detected in order to identify a lensed GRB source; thus, the overall BATSE burst detection efficiency, $\\epsilon$, is of prime importance in determining the observed incidence of lensing. The average efficiency for the 4B catalog is $\\epsilon = 0.48$ (\\cite{M98}), which is 40\\% larger than the previously estimated value (\\cite{F94}). This increase implies that the expected incidence of lensing is significantly higher than was previously believed. Lensing of GRBs was first suggested by Paczynski (1986) as a way to establish their cosmological origin. Subsequently, many authors have calculated the lensing rate of GRBs (\\cite{M92}; \\cite{BW92}; \\cite{N93}; \\cite{GN94}), assuming GRB redshifts estimated from no--evolution fits to the log N -- log P distribution, viz., $z\\lta 1$. For redshifts $z \\lta 1$, the gravitational lensing rate is reasonably well known for a given cosmology (e.g., \\cite{TOG84}; \\cite{FT91}; \\cite{F92}). At redshifts $z \\gta 1$, however, the lensing rate has been uncertain, mainly because of ambiguity in the angular diameter distance at high redshift (\\cite{F92}). As a result, even when the cosmology and the number density and properties of lenses are fixed, the estimated lensing rate can vary by factors of several between the different prescriptions. Motivated in part by this ambiguity, Holz \\& Wald (1998) have developed a numerical method to calculate lensing rates for a given cosmology, combining techniques from both ``ray-shooting'' and ``Swiss-cheese'' models. The approach resolves the angular-diameter distance uncertainties, and also correctly accounts for multiple lens encounters along the line of sight, allowing for an unambiguous calculation of lensing rates. If there are no multiple images among the bursts detected with BATSE, then an upper limit to the average redshift $\\langle z\\rangle$ of GRBs can be inferred, one which depends directly on the cosmological parameters (see also Nemiroff et al.\\ 1994; Marani et al.\\ 1998; Marani 1998). Here, we set an upper limit to $\\langle z\\rangle$, independent of the physical model for GRBs, using the Holz \\&~Wald (1998) method to compute the lensing rate for a variety of cosmologies. The plan of this paper is as follows. In \\S~2 we describe and develop the statistical lensing method, and compare its results with analytical estimates. We show that analytical estimates using the ``filled-beam\" approach are reasonably accurate for source redshifts less than $\\sim$2--3, but underestimate the true rate of lensing by tens of percent at higher redshifts. In \\S~3 we use the numerical method to compute the lensing rate as a function of redshift for several cosmologies, and compare these results with previous estimates. We also determine upper limits on the rate of lensing and on the average redshift $\\langle z\\rangle$ of GRBs, assuming that no lensing events are present in the BATSE 4B catalog (\\cite{M98}). We discuss the implications of these results and give our conclusions in \\S~4. We follow the conventions that the Hubble constant is 100\\,$h$\\ km~s$^{-1}$~Mpc$^{-1}$, $\\Omega$ is the present mean density in the universe in units of the closure density, and $\\Lambda$ is the present normalized cosmological constant. In a flat universe, $\\Omega+\\Lambda=1$. ", "conclusions": "We have exhibited a new method for calculating gravitational lensing rates, given a cosmology and a distribution and evolution of the lenses. This method, which is described more fully in Holz \\& Wald (1998), is free of the angular diameter distance ambiguity that is a feature of standard analytical methods. For the cosmologies and redshift ranges we have examined, we find that the lensing rate is approximately equal to the filled-beam lensing rate for $z<3$, but in excess of the filled-beam rate for $z>3$. We find that the lack of detected lensing in the current BATSE catalog places an upper limit to the median redshift of bursts that, at the 68\\% confidence level, is comparable to or less than the median redshift in star-formation burst models. We now discuss these results and place them in context. In \\S~4.1 we discuss the simplifying assumptions that have been made in our calculation. We find that these assumptions have, if anything, caused us to {\\it underestimate} the lensing rate. In \\S~4.2 we examine the uncertainties in the inputs used in our calculations, as well as the future prospects for reducing these uncertainties. In \\S~4.3 we discuss the implications of the current lack of lensing for cosmological models of gamma-ray bursts, when combined with existing lower bounds to the redshift of bursts in these models. Finally, in \\S~4.4 we provide a future outlook for the rapidly emerging importance of constraints inferred from detection or nondetection of lensing. \\subsection{Effects of Simplifying Assumptions} We have assumed that the angular distribution of the sources of gamma-ray bursts is uncorrelated with the angular distribution of the lenses. This is well justified, because for sources at low redshift the lenses that contribute most to the lensing rate are at about half the redshift to the source, and for sources at high redshift the dominant lenses are at a redshift of approximately unity (see \\S~2; see also, e.g., Turner et al.\\ 1984; \\cite{M92}). Hence, in a cosmological model the sources of gamma-ray bursts are separated from lenses by hundreds of megaparsecs to gigaparsecs, and therefore it is extremely unlikely that the sources and lenses are angularly correlated. We have also assumed that the phase of the orbit of BATSE is uncorrelated between the separate images of the burst, and hence that the joint probability of two images being observable is just the square of the BATSE efficiency, $0.48^2 \\approx 0.23$. This assumption is reasonable if the time delay between images is significantly larger than the BATSE orbit time of $\\sim 5000$ seconds. The time delay for a mass $M$ is of order $10^{-5}(M/M_\\odot)$ seconds (see, e.g., Blandford \\& Narayan 1992), and hence the assumption of uncorrelated phases is good for masses $M\\gta 10^9\\,M_\\odot$. This includes almost all the effective lensing mass of galaxies, and thus this assumption is also unlikely to affect the calculated lensing rates significantly. We may have underestimated the lensing rate by assuming that only galaxies---which we have modeled as singular isothermal spheres---contribute to the lensing rate. Nemiroff et al.\\ (1993) have noted that point masses, for example supermassive black holes, could also in principle contribute. To produce gamma-ray burst lensing of the type that we consider here, in which the lensing event appears as two separate bursts, the mass of such a lensing black hole must be $M\\gta 10^8\\,M_\\odot$, because smaller masses would produce time delays less than $\\sim$1000 seconds, so that overwriting or readout time would tend to prevent BATSE from detecting two separate bursts (the effect of lensing by lower-mass black holes may, however, be detectable using techniques such as autocorrelation analysis; see Nemiroff et al.\\ 1994, 1998). If, however, the total lensing rate were dominated by very massive black holes, $M\\gta 10^{10}\\,M_\\odot$, then these objects would also produce detectable image separation of lensed quasars or galaxies (the angular separation is of order $3(M/M_\\odot)^{1/2} (D/1 {\\rm Gpc})^{-1/2}\\,\\mu$arcsec for an Einstein ring, which is 0.3\" for $M=10^{10}\\,M_\\odot$ and $D$=1 Gpc). Hence, the amount by which we have underestimated the lensing rate depends on the number of point masses in the relatively narrow range $10^8$--$10^{10}\\,M_\\odot$, which is the only mass range that could avoid detection in quasar lensing surveys and yet produce separately detected bursts. We may also have underestimated the lensing rate by only keeping track of the total number of images, not whether there are, e.g., two or four images in a particular event. As pointed out by Grossman \\& Nowak (1994), this tends to underestimate the {\\it detectable} rate of lensing because when there are several images it is more probable that at least two of them are observed by BATSE: assuming a 48\\% efficiency for any particular image, the probability of observing at least one pair rises from 23\\% when there are two images to 66\\% when there are four images. Grossman \\& Nowak (1994) estimate that, for galaxy ellipticities typical of galaxy surveys, the maximum possible enhancement of the lensing detection rate is $\\sim$30\\%, but that the overall enhancement is likely to be smaller. A third effect that may enhance the rate of lensing is magnification bias, which is an effect first discussed extensively by Fukugita \\& Turner (1991) in the context of quasar lensing. Magnification bias occurs because sources that would have been undetectable are made visible by lensing, and hence a magnitude-limited sample contains an enhanced incidence of lensing. This can be especially important if the number of sources at a given flux rises steeply at the faint end, as it does in quasars. To detect multiple images and thus confirm strong lensing, it is necessary to detect the fainter image as well as the stronger image. For persistent sources such as quasars, this can be done by following up broad, shallow surveys by deep pointings, so that essentially all of the secondary images are detectable (e.g., as in the Hubble Snapshot Survey [Bahcall et al. 1992; Maoz \\etal\\ 1992, 1993a,b]). For transient sources such as gamma-ray bursts, deep follow-ups are not always possible, and hence for the secondary image to be detectable it must have a brightness in excess of the threshold of the original survey (i.e., in the case of gamma-ray bursts, the BATSE sample). This effect, plus the flatness of the faint end of the gamma-ray burst log N -- log P curve, suggests that the magnification bias for gamma-ray bursts is probably less than the magnification bias for quasars (see also \\cite{GN94} for a discussion of this point). Nonetheless, magnification bias for gamma-ray bursts could in principle increase significantly the expected lensing rate, and hence decrease the upper limits on $\\langle z\\rangle$, compared to the conservative estimates here. We may underestimate the lensing rate because we have neglected the clustering of galaxies, the last of our assumptions. The enhanced gravitational potential in clusters tends to increase the convergence of null geodesics and hence increase the cross section for strong lensing (see also Holz \\& Wald 1998). Numerical estimates suggest that the lensing rate for galaxies in a cluster will be increased by $\\sim$10--20\\% by this effect, but because only $\\sim$10\\% of galaxies are in clusters this effect is unlikely to increase the overall lensing rate significantly. Finally, however, we may have overestimated the rate of lensing by ignoring evolution in the properties of the lensing galaxies. If the comoving density in lensing galaxies was less in the past than it is now, or if these galaxies were less massive than they are today, high-redshift lenses would contribute less to the lensing rate than we have assumed. Note, however, that the peak in the lens redshift distribution is at about unity (Fig.~2), when galaxies had properties very similar to their current properties. Note also that observations of lensing of quasars (see below), which have a typical redshift similar to the proposed typical redshift $z\\sim 2$ of GRBs, suggest that if there are evolutionary effects then these effects have not had a dominant effect on the lensing rate. To summarize, the net effect of our simplifying assumptions is likely to be that we underestimated the lensing rate by $\\lta$10\\%, and hence our results our conservative in that they give a slightly high upper bound to $\\langle z\\rangle$. \\subsection{Uncertainties in Input Parameters} In addition to the simplifying assumptions discussed above, our calculations are clearly dependent on the input values we have assumed. One input is the dimensionless parameter $F$, which is proportional to the product of the number density of galaxies and the fourth power of their average velocity dispersion. The lensing rate scales linearly with $F$. We have used $F$=0.1, which is consistent with the recent estimates from the Century Survey (Geller et al.\\ 1997). Other estimates range from $F$=0.05 to $F$=0.15, and if these other estimates are used then the range of possible maximum redshifts is increased for a given cosmology. Fundamental cosmological parameters, another input to our calculations, are also uncertain. A standard Einstein-De Sitter universe ($\\Omega$=1, $\\Lambda$=0) gives the lowest lensing rate, and a $\\Lambda$-dominated universe gives the highest. Current observation of quasar lensing puts constraints on the allowed combinations of $F$ and cosmology. For example, five of the 351 quasars in the HST Snapshot Survey (Maoz et al.\\ 1993b) were lensed. A comparison of this rate with the rate expected for models is complicated by effects such as magnification bias (see above). However, the high average redshift of the quasars ($z\\sim 2$) means that many of the issues affecting lensing of gamma-ray bursts, such as evolution of clustering, also affect the lensing of quasars. Hence, quasar lensing statistics have bearing on estimated GRB lensing rates, and will become even more relevant as more extensive surveys are done. In addition, all of these uncertainties will be diminished greatly by the data that emerge from the plethora of satellites and surveys planned for the next few years. The number density and velocity distribution of galaxies (and hence $F$) will be determined with unprecedented accuracy by surveys such as 2dF (Colless 1998) and the Sloan Digital Sky Survey (Margon 1998), the latter of which is expected to see first light in 1998. Data from these surveys will also provide valuable information about $\\Omega$, $\\Lambda$, clustering, and the lensing rate of quasars. Complementary information about cosmological parameters will be extracted from high-redshift supernova surveys (Perlmutter \\etal\\ 1998) and from data gathered by the many upcoming microwave background experiments, such as MAP (Wang, Spergel, \\& Strauss 1998), TOPHAT (Martin et al.\\ 1996), and PLANCK (see, e.g., Bond, Efstathiou, \\& Tegmark 1997 for a discussion of the constraints). The evolution of galaxy clustering is already being inferred from cluster surveys, and AXAF observations of X-ray emission from clusters of galaxies is expected to greatly enhance this understanding. Hence, one side effect of the coming data-rich era in cosmology is that calculations of lensing rates will be much less uncertain, and therefore upper limits to the redshift of a population such as gamma-ray bursts will be far more secure. \\subsection{Current Implications for Cosmological GRB Models} The redshift limits presented in this paper have important consequences for cosmological models of GRBs. Lower limits on the redshift of bright GRBs are becoming stronger as the number of detected bursts rises, both because of the no-host problem (Schaefer et al.\\ 1997) and because GRBs do not show any evidence of large-scale clustering (Lamb \\& Quashnock 1993; Quashnock 1996). These limits suggest that if the sources of GRBs are in or near galaxies, the median redshift of the bursts detected with BATSE may be significantly greater than unity. However, the upper limits to the redshift of GRBs that follow from the lack of lensing are beginning to make such high-redshift populations less appealing. For example, the lack of lensing in the current BATSE catalog strongly rules out the \\hbox{$z\\sim 10$--200} models that were heretofore consistent with the properties of the BATSE bursts (Rutledge et al.\\ 1995). Also, models in which the burst rate is proportional to the star formation rate (Totani 1997; Wijers et al.\\ 1998), and for which $\\langle z\\rangle\\sim 2.2$, are beginning to be constrained strongly by the lack of lensing: Our 68\\% upper limit on $\\langle z\\rangle$ (eq.~[6]) is less than 1.9, and probably close to 1.0. If the redshift upper limits are reduced further, it may be necessary to postulate either a very short interval in which GRBs were produced, with an intrinsic brightness distribution that matches the observed log N -- log S distribution, or a population that is unassociated with any known population of objects, so that the lower limits to redshift do not apply. \\subsection{Future Outlook and Summary} In the next few years gravitational lensing is likely to play a role of rapidly increasing importance in constraining cosmological models of gamma-ray bursts. As discussed in \\S~4.2, the uncertainties in the calculation of the expected lensing rate will be diminished greatly, and hence the limits will be robust. In addition, as BATSE continues to detect bursts the statistics will steadily improve, and in the current high-redshift models lensing is to be expected in the next few years. In five years, the sample will be roughly double the $\\sim$1800 bursts considered here, and therefore if lensing continues to be absent, upper limits on the redshift will become extremely restrictive. At the same time, the improved statistics will continue to increase the lower limits on redshift, if there is no apparent large-scale clustering (Lamb \\& Quashnock 1993; Quashnock 1996). Moreover, the accurate positional estimates of BeppoSAX, HETE II, and the Fourth Interplanetary Network (IPN) are likely to provide $\\sim$100 error boxes of area a few square arcminutes or less in the next five years. If instead lensing {\\it is} detected, there will be many important consequences. For one, the cosmological origin of a particular GRB will be established from gamma-ray data for the first time (as opposed to from the afterglow, as may be the case for GRB~970508 [\\cite{M97}]). The approximate mass of the lens, and hence its luminosity, can be estimated from the time delay between images. The sharp peak in the probable redshift of lenses (see \\S~2) would then give a reasonably accurate estimate of the flux of the lens. If the GRB can be reasonably well-localized (e.g., with the IPN), this will allow a search for the lens. If the lens is found, the GRB source must essentially be angularly coincident with it, because the deflection angle is only about 1\", and thus counterpart searches will be facilitated greatly. Knowledge of the redshift of the lens, combined with $H_0$ and an independent estimate of the mass of the lens (e.g., from its luminosity or velocity dispersion) will allow use of the time delay to estimate the redshift of the source. The quantifiable diversity of gamma-ray burst light curves (\\cite{MN98}; \\cite{Mar98}) means that false positives are unlikely, and hence even a single lensing event will generate a wealth of data. In conclusion, the increased estimates of the redshift of gamma-ray bursts that were forced by the no-host problem, in addition to the greatly improved statistics of bursts and the increase in the estimate of the BATSE burst detection efficiency from 34\\% to 48\\%, mean that gravitational lensing limits are now playing a prominent role in constraining gamma-ray burst models. The role of lensing will become dramatically more important and robust in the next five years." }, "9804/astro-ph9804101_arXiv.txt": { "abstract": "The blazar Mkn 421 has been observed, as part of the AO1 Core Program, five times from 2 to 7 May 1997. In the LECS+MECS energy band the spectrum shows convex curvature, well represented by a broken power--law. Flux variability (more than a factor 2) has been detected over the entire 0.1--10 keV range, accompanying which the spectrum steepens with the decrease in intensity. Mkn 421 has also been detected with the PDS instrument. Our preliminary analysis indicates that the PDS spectrum lies significantly above the extrapolation from the MECS, suggesting a contribution from a flatter high energy component. ", "introduction": "Mkn 421 is one of the best known and studied BL Lac objects. It shows optical polarization, flat radio spectrum, and large variability, characteristics of the blazar class. It is bright in X--rays, where it shows prominent flares accompanied by significant spectral changes \\cite{takahashi}. Among GeV--emitting blazars, Mkn 421 is unique in being the first object in which the $\\gamma$--ray emission is detected to extend up to TeV energies (\\cite{punch,krennrich}), at a level allowing detailed spectral and variability studies in the broadest available range of frequencies. Its GeV (EGRET) $\\gamma$--ray emission connects smoothly with the E$>$0.5 TeV spectrum (e.g.~\\cite{macomb}). On the contrary the X--ray spectrum, up to 10 keV, did not show any hint of the onset of the inverse Compton component responsible for GeV and TeV emission. It was then of great interest to take advantage of the full capabilities of {\\it Beppo}SAX, enabling a spectral coverage up to $\\gta$ 200 keV, to look for it. At the same time the fact that in the X--ray band we are observing the emission from the highest energy tail of the emitting--particle distribution could provide important clues on particle acceleration and cooling mechanisms. The Mkn~421 observations discussed here are part of the AO1 Core Program dedicated to bright blazars. The data reduction presented here has been generally performed with software released {\\it before} September 1997. Data reduction with the updated software for all the on board instruments, and a more detailed analysis, will be presented in forthcoming paper with all appropriate references. \\begin{figure}[t] \\vspace{9pt} \\psfig{file=lecs_2mecs_lc.ps,width=7.5truecm,rheight=5.3truecm} \\caption{\\small\\sf Rebinned light curve (4000~s bins) of {\\it Beppo}SAX data, LECS, and MECS divided in two energy bands [1.5 -- 4], [4 -- 10] keV.} \\label{fig:light_curve} \\end{figure} \\begin{table*}[t] \\setlength{\\tabcolsep}{0.7pc} \\newlength{\\digitwidth} \\settowidth{\\digitwidth}{\\rm 0} \\catcode`?=\\active \\def?{\\kern\\digitwidth} \\caption{{\\it Beppo}SAX Observations Log} \\label{tab:log} \\begin{tabular*}{\\textwidth}{@{}l@{\\extracolsep{\\fill}}ccccc} \\hline {Pointing} & {LECS } & {LECS } & {MECS } & {MECS } & {MECS } \\\\ {Start Date} & {exp. time} & {[0.1--4 keV]} & {exp. time} & {[1.5--10 keV]} & {[$<$4/$>$4 keV]} \\\\ { } & {(ksec)} & {(cts/s)} & {(ksec)} & {(cts/s)} & {(cts/s)} \\\\ \\hline { 2/V/1997 @ 04:10 } & { 4.4 } & {$ 2.806 \\pm 0.020 $} & { 11.4 } & {$ 2.862 \\pm 0.016 $} & { (2.14/0.72) } \\\\ { 3/V/1997 @ 03:24 } & { 4.3 } & {$ 1.748 \\pm 0.020 $} & { 11.7 } & {$ 1.694 \\pm 0.012 $} & { (1.28/0.41) } \\\\ { 4/V/1997 @ 03:25 } & { 4.9 } & {$ 1.362 \\pm 0.017 $} & { 12.2 } & {$ 1.027 \\pm 0.009 $} & { (0.81/0.22) } \\\\ { 5/V/1997 @ 03:32 } & { 4.9 } & {$ 1.823 \\pm 0.029 $} & { 11.9 } & {$ 1.522 \\pm 0.012 $} & { (1.52/0.35) } \\\\ { 7/V/1997 @ 04:47 } & { 6.0 } & {$ 1.612 \\pm 0.016 $} & { \\nodata } & { \\nodata } & { \\nodata } \\\\ \\hline \\end{tabular*} \\end{table*} ", "conclusions": "{\\it Beppo}SAX dat aof Mkn 421 shows interesting variability both in flux and in spectral shape, with a marked softening corresponding to decreasing brightness. This kind of spectral variability behavior is well known in the X--ray band for sources of the class of Mkn~421, the so called High-Frequency-Peaked BL Lacs (HBL). In general in blazars at energies just above the synchrotron peak it the relationship {\\it harder-when-brighter} holds and is generally interpreted in terms of injection of fresh electrons in the highest energy end of a single population. On May 2$^{\\rm nd}$ the peak of the synchrotron component could possibly fall in/or just below the LECS band, while in the lower state of May 4$^{\\rm th}$ this is not longer true. Moreover, preliminary analysis of PDS data suggests the presence of a deviation from the continuously downward curvature for E $> 10$ keV, possibly being the signature of the onset of a different harder spectral component. Further temporal and spectral analysis, also with comparison of multifrequency data and comparison with model predictions is in progress." }, "9804/astro-ph9804337_arXiv.txt": { "abstract": "A statistical study of global galaxy parameters can help to improve our understanding of galaxy formation processes. In this paper we present the analysis of global galaxy parameters based on optical and near-infrared observations of a large sample of edge-on disc galaxies. \\\\ We found a correlation between the ratio of the radial to vertical scale parameter and galaxy type: galaxies become systematically thinner when going from S0's to Sc's, whereas the distribution seems to level off for later types. \\\\ The observed scale length ratios (and thus the radial colour gradients) largely represent the galaxies' dust content. On average the colour gradients indicated by the scale length ratios increase from type Sa to at least type Sc. For galaxy types later than Sc, the average colour gradient seems to decrease again. \\\\ The distribution of {\\it K}-band (edge-on) disc central surface brightnesses is rather flat, although with a large scatter. However, the latest-type sample galaxies ($T > 6$) show an indication that their average disc central surface brightnesses may be fainter than those of the earlier types. This effect is probably not the result of dust extinction. ", "introduction": "A study of the statistical properties of highly inclined, or ``edge-on'' galaxies benefits greatly from the special orientation with respect to the line of sight of such galaxies. Observations of edge-on galaxies provide us with direct measurements of the luminosity and colour distributions both perpendicular to the galaxy planes and along the galaxies' major axes at various heights above the plane. Indirectly, these luminosity distributions can be related to the galaxies' density distributions and thus their global structure. Moreover, in-depth knowledge of the dust distribution, and hence the optical depth of galaxies, is important for our understanding of galaxy evolution. \\subsection{The flattening of exponential discs} \\label{ratio.sec} A major advantage of studying highly inclined galaxies is that one can determine their radial and vertical scale parameters directly and independently, since the dependence of these parameters on inclination is smallest for the highest inclinations (e.g., van der Kruit \\& Searle 1981a; de Grijs et al. 1997). These scale parameters provide us with information about the intrinsic shape of galaxy discs, i.e., their flattening, in a more direct way than the canonical axis ratios. Moreover, since the vertical scale height, $z_0 = 2 h_z$ (where $h_z$ is the exponential scale height), is to first order independent of radius (e.g., van der Kruit \\& Searle 1981a,b, 1982a; Kylafis \\& Bahcall 1987; Shaw \\& Gilmore 1990; Barnaby \\& Thronson, Jr. 1992; but see de Grijs \\& Peletier 1997), the radial to vertical scale parameter ratio, $h_R/z_0$, can often be determined more accurately than the major to minor axis ratio. By studying the scale parameter ratio statistically, we may be able to put constraints on the disc formation processes as well as on the stability of galaxy discs (e.g., Bottema 1993). When considering the physical processes that determine the scale parameters one does not immediately expect a strong correlation between scale length and scale height. The scale height is likely determined by the internal, secular evolution of the stellar velocity dispersion (e.g., van der Kruit \\& Searle 1981a; Carlberg 1987), whereas the scale length is basically the result of the composition of the protogalaxy (Fall 1983; van der Kruit 1987). However, one might expect that in a larger disc, with a greater rotation velocity, the heating of the disc stars may be more violent, thus resulting in a larger scale height. Therefore, one may expect a correlation between the rotation velocity (which can be related directly to the scale length) of a galaxy disc and the scale height, although the precise dependence is yet unknown (see, e.g., Bottema 1993). Thus, statistics on the ratio of scale length to scale height can be expected to give information on the importance of the formation processes in disc galaxies with different properties. Moreover, once the $h_R/z_0$ ratio is known, one may be able to determine the (theoretical) maximum rotation of a disc from measurements of the vertical disc dispersion (Bottema 1993). Therefore, a statistical treatment of the scale parameter ratio may put general constraints on both the kinematical properties and the global stability of galaxy discs. Bottema (1993) predicts that a constant value for the $h_R/z_0$ ratio leads to a more or less constant mass-to-light ratio of the old disc, $(M/L)_B$, under the assumption that we are dealing with exponential, locally isothermal discs with a constant ratio of vertical to radial velocity dispersion. On the other hand, if we assume a linear relationship between the old-disc absolute luminosity and the vertical velocity dispersion, Bottema (1993) shows that, for a constant $(M/L)_B$, the $h_R/z_0$ ratio decreases rapidly from faint galaxies to a constant level for normal and bright galaxies. Thus, in general, the observed velocity dispersions imply that a constant old-disc mass-to-light ratio results in an approximately constant scale parameter ratio, whereas a constant scale parameter ratio also leads to a mass-to-light ratio that is, to first order, constant. In fact, these predictions imply that all galaxy discs are governed by equal mass-to-light ratios in the old stellar populations, assuming that all galaxy discs have approximately the same colour (Bottema 1993). However, the assumption of a constant and equal mass-to-light ratio of the old-disc population in disc galaxies is probably not physically realistic, considering the range of colours observed within and among galaxies (e.g., de Jong 1996c, and references therein). Therefore, the predicted relationships should be treated with caution and only be used as general guidelines. \\subsection{Colour gradients as diagnostics} Broad-band colours are relatively easy to obtain and are therefore the most widely used colour diagnostics to date. They immediately reveal the approximate nature of a galaxy, which is to first order determined by its dominant stellar population and dust content. Although for the detailed analysis of galaxy luminosity and colour profiles one needs to adopt {\\it a priori} assumptions concerning the evolutionary stellar population synthesis, the initial mass function, the metallicity and the star formation history, as well as about the dust geometry and its characteristics, de Jong (1996c) shows that the colours formed from different broad-band combinations correlate strongly, which indicates that these colours are probably caused by the same physical process. Therefore, broad-band colours can be used as indicators of changes in the gross properties of galaxies (e.g., changes in metallicity and/or dust contamination). All systematic colour differences induced by stellar population changes and metallicity gradients are generally considerably smaller than the reddening due to dust, however. \\subsubsection{Radial colour gradients in edge-on disc galaxies} \\label{edgeongrad.sect} In contrast to the large number of studies of radial colour gradients in moderately inclined and face-on spiral galaxies (e.g., de Jong 1996c, and references therein), the colour behaviour of highly inclined and edge-on galaxies has not received much attention. In highly inclined galaxies, the study and interpretation of intrinsic colour gradients is severely hampered by the presence of dust in the galaxy planes, which causes the dust lane to appear as a red feature in vertical colour profiles (e.g., Hamabe et al. 1979; Hegyi \\& Gerber 1979; van der Kruit \\& Searle 1981b; Jensen \\& Thuan 1982; de Grijs et al. 1997). In individual edge-on galaxies, it is generally found that the colours along the major axes, i.e., the locations of the dust lanes, remain nearly constant (e.g., Sasaki 1987; Wainscoat et al. 1990; Aoki et al. 1991; Peletier \\& Balcells 1997), although in most cases the outermost disc regions tend to be slightly bluer on the major axis (e.g., Sasaki 1987), which may be explained in terms of an increasingly metal-poor population or a decreased amount of dust at larger galactocentric distances. Generally, as the height above the dust lane and its embedded young disc increases, the radial colour gradients become small or statistically insignificant (e.g., Hamabe et al. 1979, 1980; van der Kruit \\& Searle 1982a,b; Jensen \\& Thuan 1982; Peletier \\& Balcells 1997). \\subsubsection{Colour gradients from scale length ratios} \\label{colgrads.sect} Since the dust influence varies as a function of passband, scale length ratios could be used as a diagnostic to estimate colour gradients and the dust content of a given galaxy. Evans (1994) studied the effects of dust on the stellar scale length as a function of wavelength, under the assumption that the resulting scale length differences are solely due to dust absorption. His models predict that these differences are small, at least for face-on galaxies, on the order of the observational uncertainties, and even smaller for galaxies with a prominent bulge component. According to his models, if the scale height ratio between dust and stars is $\\sim 0.5$ (Peletier \\& Willner 1992; Evans 1994), Evans' (1994) models exclude face-on galaxies with $h_B/h_H \\approx 2$. On the other hand, larger ratios can be obtained if a galaxy is inclined with respect to the line of sight. The measurement of blue to red scale length ratios alone will not unambiguously reveal the dust content of a given galaxy, because any deviation from unity can equally well be explained by an intrinsic colour gradient, especially for face-on galaxies (Byun et al. 1994). ", "conclusions": "" }, "9804/astro-ph9804011_arXiv.txt": { "abstract": "In this paper we demonstrate for the first time the connection between the spatial and temporal progression of star formation and the changing locations of the very dense regions in the gas of a massive disk galaxy (NGC~1144) in the aftermath of its collision with a massive elliptical (NGC~1143). These two galaxies form the combined object Arp 118, a collisional ring galaxy system. The results of 3D, time-dependent, numerical simulations of the behavior of the gas, stars, and dark matter of a disk galaxy and the stars and dark matter in an elliptical during a collision are compared with multiwavelength observations of Arp 118. The collision that took place approximately 22 Myr ago generated a strong, non-linear density wave in the stars and gas in the disk of NGC~1144, causing the gas to became clumped on a large scale. This wave produced a series of superstarclusters along arcs and rings that emanate from the central point of impact in the disk. The locations of these star forming regions match those of the regions of increased gas density predicted the time sequence of models. The models also predict the large velocity gradients observed across the disk of NGC 1144. These are due to the rapid radial outflow of gas coupled to large azimuthal velocities in the expanding ring, caused by the impact of the massive intruder. ", "introduction": "Optically-identified, collision-produced 'ring galaxies' usually display vigorous star formation in an expanding ring or arc of high density gas (e.g., Joy \\& Harvey, 1987; Joy \\& Ghigo, 1988; Appleton \\& Marston, 1997; see Appleton \\& Struck-Marcell, 1996). This recent star formation often dominates the appearance of these galaxies, with H$\\alpha$ emission tracing the giant HII regions (e.g., Hippelein, 1989a) and outshining the nucleus unless the latter is active. Near-infrared images of these systems show that the older stellar disk population has also been swept into an outwardly expanding, strong density wave. The density disturbance produced by a collision through (and roughly perpendicular to) a galaxy's disk propagates outward through a rotating disk of gas and stars, which is itself at first contracting and then expanding as the collision proceeds. This superposition of material motions and wave propagation produces a pattern of both closed loops and open-ended arcs of relatively high density gas. Shocks can occur in these regions because of the high relative velocities that are produced in the flows. The higher density features are well delineated in observations of these systems and in the combined N-body/hydrodynamic models that we and others have produced. The models can be exploited to further our understanding of the star formation that is triggered by a galaxy collision, a process that has likely occurred in many galaxies over their lifetimes and may have been very frequent earlier in the universe (see Lavery {\\it et al.}, 1996). Here we compare our numerical models of these types of galaxy systems with observational data on Arp 118, one particular IR-luminous, gas-rich example of a collisionally produced ring galaxy. This system consists of a strongly distorted disk galaxy and an elliptical in close proximity. Hippelein (1989a) had difficulty in explaining the extreme velocity gradient and complex morphology in the Arp 118 system using a his simple picture of a collision between a gas-rich spiral and an elliptical. However, Gao {\\it et al.} (1997) have observed that Arp 118 contains a large amount of molecular gas distributed exclusively along the ring formations (the first high resolution CO observations ever made of a ring galaxy) , and that the velocity structure in this gas is kinematically consistent with the simple collisional model. In this paper, we show that, using fully dynamical, 3D models, we can reproduce the morphology of the disturbed disk galaxy in the pair and the approximate relative positions of the two galaxies. The 'best-fit' model for the Arp 118 system was chosen from a grid of simulations produced by Gerber, Lamb, \\& Balsara (1996). These simulations explore the results of face-on collisions (collisions parallel to the spin-axis of the disk) between an elliptical galaxy and a disk galaxy. This dynamical treatment confirms the correspondence between the models and the observed velocity structure. We use the chosen simulation to constrain the timescales for star formation in Arp 118 and to explore the history of the collision by comparing the predicted results of such a collision with current observations. ", "conclusions": "The simulation presented here is comprised of a sequence of models that follows the evolution of a pair of galaxies through and after a collision. One of the later models closely matches the observed morphology of the CO component in Arp 118 and its velocity field. Earlier models in the sequence provide a good match with the present radio continuum and H$\\alpha$ emission, indicating that the sequence of post-collision star formation can be traced and timed by comparison to simulations of encounters between two massive galaxies with masses similar to those observed. Stars, particularly massive stars, are formed in the cores of giant molecular clouds in the highest density regions. Both multiwavelength observations and our numerical models of Arp 118 indicate that strong shocks in the gas together with large increases in the gas volume density are associated with star formation over volumes of 1 kpc$^3$. The observed morphology of the regions of dense gas and the clustering observed in the stars emitting H$\\alpha$, which were formed in the gas since the collision, suggest that the observed clumping of the young stars results from a clumping of the densest gas on the same scale. The recent work by Marston \\& Appleton (1995) and Appleton \\& Marston (1997) also provides evidence that the clumping observed in the optical images of collisionally produced ring galaxies is not due to patchy dust obscuration, because the same clustering is also observed in the near infrared. Gas clumping on this same scale is found in the numerical simulations, suggesting that there is a global explanation for the observed morphology of the dense gas and the resulting giant stellar formations in these systems. The simulations show a relatively small perturbation (clumping) in the density of stars. We therefore predict that the distribution of the old stars in these systems will be observed to be smoother than that of the gas, although the models show that the stars are driven into a wide ring, and sometimes even a second, inner ring, by the collision (see Lamb {\\it et al.} 1998). This study of the Arp 118 system demonstrates that a careful comparison between high resolution observations and detailed models can yield insight into the sequence of star formation that takes place in a gas-rich galaxy after a major collision. The intensity and location of the starburst at any particular epoch will depend upon the speed with which density waves are propagating through the expanding disk. Such quantities can now be predicted quite accurately from current models of colliding galaxies. Thus global star formation (on the scale of several hundred parsecs) as it occurs in these systems at the current epoch can be investigated more thoroughly than previously. The rate of galaxy collisions in the past was larger than it is today, due to the greater overall density, thus a considerable portion of the star formation that took place in young disk galaxies at earlier epochs was likely triggered by galaxy collisions. We expect, therefore, that studies like the one reported here will help in understanding this earlier star formation and its current consequences." }, "9804/astro-ph9804227_arXiv.txt": { "abstract": "A comparison is made between the properties of CAL 83, CAL 87, RX J0513.9$-$6951, 1E 0035.4$-$7230 (SMC 13), RX J0019.8+2156, and RX J0925.7$-$4758, all supersoft X-ray binaries. Spectra with the same resolution and wavelength coverage of these systems are compared and contrasted. Some new photometry is also presented. The equivalent widths of the principal emission lines of H and He II differ by more than an order of magnitude among these sources, although those of the highest ionization lines (e.g. O VI) are very similar. In individual systems, the velocity curves derived from various ions often differ in phasing and amplitude, but those whose phasing is consistent with the light curves (implying the lines are formed near the compact star) give masses of $\\sim$1.2M$_{\\odot}$ and $\\sim$0.5M$_{\\odot}$ for the degenerate and mass-losing stars, respectively. This finding is in conflict with currently prevailing theoretical models for supersoft binaries. The three highest luminosity sources show evidence of ``jet\" outflows, with velocities of $\\sim$1--4$\\times$10$^3$ km s$^{-1}$. In CAL 83 the shape of the He II 4686\\AA\\ profile continues to show evidence that these jets may precess with a period of $\\sim$69 days. ", "introduction": "The close-binary ``supersoft sources\" (SSS) are now recognized as a distinct class of very luminous (L$_{bol}\\geq10^{38}$ erg s$^{-1}$) X-ray sources characterized by extremely soft X-ray spectra with little or no radiation above $\\sim$0.5 keV (e.g. Tr\\\"umper et al.\\ 1991). Several reviews of the observational properties of these sources have recently been published (e.g. Hasinger 1996, Greiner 1996, Kahabka \\& van den Heuvel 1997). All SSS appear to have high mass-accretion rates and exhibit long-term X-ray and optical variability which are thought to reflect variations in the rate of mass transfer. In addition, some SSS show evidence of collimated outflows or ``jets\" (Crampton et al.\\ 1996, hereafter CHC96; Southwell, Livio \\& Pringle 1997). Van den Heuvel et al.\\ (1992) suggested that the X-ray properties of SSS are best explained by a model involving steady nuclear burning on the surface of a white dwarf accreting material at the Eddington rate. Many observations appear to support this model (Greiner 1996), although alternative interpretations (e.g. Kylafis \\& Xilouris 1993) have not yet been ruled out. During a 1996 November CTIO observing run we obtained spectra and some photometry of six close-binary supersoft sources. One of these lies in the Small Magellanic Cloud, 1E 0035.4$-$7230 (hereafter SMC 13). CAL 83, CAL 87, and RX J0513.9$-$6951 (hereafter RX J0513) are all members of the Large Magellanic Cloud, while RX J0019.8+2156 (hereafter RX J0019) and RX J0925.7$-$4758 (hereafter RX J0925) are galactic systems. Since these objects were observed with the same spectrographic configuration, intercomparison of their spectra is very straightforward. In addition, we present previously unpublished photometry for CAL 83 and a few observations of RX J0513 and RX J0925. New data for CAL 87 is being published in a separate paper (Hutchings et al.\\ 1998). Long-term spectroscopic and photometric monitoring of these sources is important since the SSS exhibit significant variations over timescales of months and years. A summary of the properties of these six supersoft binaries is given in Table 1, where they are listed in order of decreasing orbital period. The bolometric luminosity (L$_{bol}$) listed is the average value given in Greiner's catalog (1996); it depends strongly on the assumed model, adopted distance, and amount of absorption assumed. ", "conclusions": "" }, "9804/astro-ph9804157_arXiv.txt": { "abstract": "The scattering diameters of \\sgra\\ and several nearby OH masers ($\\approx 1\\arcsec$ at 1~GHz) indicate that a region of enhanced scattering is along the line of sight to the Galactic center. We combine radio-wave scattering data and free-free emission and absorption measurements in a likelihood analysis that constrains the following parameters of the GC scattering region: The GC-scattering region separation, $\\delgc$; the angular extent of the region, $\\psi_\\ell$ and $\\psi_b$; the outer scale on which density fluctuations occur, $l_0$; and the gas temperature, $\\te$. The maximum likelihood estimates of these parameters are $\\delgc = 133_{-80}^{+200}$~pc, $0.5\\arcdeg \\le \\psi_\\ell \\lesssim 1\\arcdeg$, and $(l_0/1\\,\\mathrm{pc})^{2/3}\\te^{-1/2} = 10^{-7 \\pm 0.8}$. The parameter $\\psi_b$ was not well constrained and we adopt $\\psi_b = 0\\fdg5$. The close correspondence between $\\delgc$ and $\\psi_\\ell\\dgc$ suggests that the scattering region encloses the \\hbox{GC}. As host media for the scattering, we consider the photoionized surface layers of molecular clouds and the interfaces between molecular clouds and the $10^7$~K ambient gas. We are unable to make an unambiguous determination, but we favor the interface model in which the scattering medium is hot ($\\te \\sim 10^6$~K) and dense ($n_{\\mathrm{e}} \\sim 10$~cm${}^{-3}$). The GC scattering region produces a 1~GHz scattering diameter for an extragalactic source of 90\\arcsec, if the region is a single screen, or 180\\arcsec, if the region wraps around the GC, as appears probable. We modify the Taylor-Cordes model for the Galactic distribution of free electrons in order to include an explicit GC component. We predict that pulsars seen through this region will have a dispersion measure of approximately $2000$~pc~cm${}^{-3}$, of which approximately 1500~pc~cm${}^{-3}$ arises from the GC component itself. We stress the uniqueness of the GC scattering region, probably resulting from the high-pressure environment in the \\hbox{GC}. ", "introduction": "\\label{sec:gc.intro} Davies, Walsh, \\& Booth~(1976) established that the observed diameter of \\sgra, the compact source in the Galactic center, scales as $\\lambda^2$, as expected if interstellar scattering from microstructure in the electron density determines the observed diameter. The observed diameter of \\sgra\\ is now known to scale as $\\lambda^2$ from 30~cm to 3~mm (\\cite{rogersetal94}) and to be anisotropic at least over the wavelength range 21~cm to 7~mm (\\cite{bzkrml93}; \\cite{krichbaumetal93}; \\cite{y-zcwmr94}). Maser spots in OH/IR stars within 25\\arcmin\\ of \\sgra\\ also show enhanced, anisotropic angular broadening (\\cite{vfcd92}; \\cite{fdcv94}). These observations indicate that a region of enhanced scattering with an angular extent of at least 25\\arcmin\\ in radius (60~pc at 8.5~kpc) is along the line of sight to \\sgra. At 1~GHz the level of angular broadening produced by this scattering region is roughly 10 times greater than that predicted by a recent model for the distribution of free electrons in the Galaxy (\\cite{tc93}, hereinafter TC93), even though this model includes a general enhancement of scattering toward the inner Galaxy. These observations do not constrain the \\emph{radial} location of the scattering region for the following reason: All previous observations have been of sources in or near the Galactic center, and for such sources, a region of moderate scattering located far from the Galactic center can produce angular broadening equivalent to that from a region of intense scattering located close to the Galactic center. Previous estimates for the location of the scattering region have ranged from~10~pc to~3~kpc. Ozernoi \\& Shisov~(1977) concluded that an ``unrealistic'' level of turbulence is implied unless the region is within 10~pc of the Galactic center. The level of turbulence they considered unrealistic, however, namely $\\sqrt{\\langle n_{\\mathrm{e}}^2\\rangle}/\\langle n_{\\mathrm{e}}\\rangle \\sim 1$, does appear to occur elsewhere in the interstellar medium (\\cite{s91}). Further, van~Langevelde et al.~(1992) used the free-free absorption toward \\sgra\\ to constrain the region's distance from the Galatic center to the range 0.85--3~kpc, though suitable adjustment of free parameters (outer scale and electron temperature) can decrease the limit to 0.03~kpc. We shall refer to the case in which the region is a site of extreme scattering, $\\lesssim 100$~pc from the Galactic center and presumably caused by processes occurring there, as the GC model. We shall refer to the case in which the region is far from the GC, $\\gtrsim 1$~kpc and a site of enhanced but not extreme scattering, as the random superposition (RS) model. Although the GC model is attractive for phenomenological reasons, other sites of enhanced interstellar scattering are found throughout the Galaxy (e.g., NGC~6634, \\cite{mrgb90}; Cyg~X-3, \\cite{mmrj95}) and the mean free path for encountering such a region is approximately 8~kpc (\\cite{cwfsr91}). Identifying the location of the scattering is important in establishing the origin of the scattering. Associating the scattering with a specific region may elucidate the mechanism for the generation of the density fluctuations responsible for the scattering. The currently favored mechanism is that velocity or magnetic field fluctuations---or both---generate the density fluctuations (\\cite{h84}, 1986; Montgomery, Brown, \\& Matthaeus~1987; \\cite{s91}; \\cite{sg94}; \\cite{gs95}). Velocity or magnetic field fluctuations are also a natural means for inducing anisotropy in the density fluctuations and thereby in the scattering disks. If this mechanism is correct, the amplitude of the density fluctuations may provide a measure of the coupling between the density and velocity or magnetic field fluctuations or, more generally, provide information about the small-scale velocity or magnetic field in the scattering region. However, current observational constraints are uncertain by the ratio of the Galactic center-scattering region distance to the Galactic center-Sun distance. In the RS model, the ratio is a few while in the GC model the ratio could be as large as one hundred, so the location of the scattering region is a key free parameter. The location of the scattering region also has implications for pulsar searches toward the \\hbox{GC}. Cordes \\& Lazio~(1997) showed that even if the RS model is correct, pulsars seen through the scattering region will suffer pulse broadening of at least 5~s at 1~GHz (see also \\cite{dwb76}; \\cite{os77}). If the GC model is correct, only at frequencies above 10~GHz will pulsations be detectable (because of the $\\nu^{-4}$ dependence of pulse broadening) and then only for pulsars with periods longer than 100~ms. In this paper we develop a likelihood analysis to quantify the most probable $\\delgc$ for the scattering region. In \\S\\ref{sec:gc.model} we describe our model for the distribution of free electrons in the \\hbox{GC}. In \\S\\ref{sec:likefunc} we assemble measurements from the literature relevant to radio-wave scattering and develop a likelihood method to constrain the properties of the scattering region, and in \\S\\ref{sec:gc.conclude} we discuss our results and present our conclusions. ", "conclusions": "\\label{sec:gc.conclude} \\subsection{Comparison with Previous Analyses} Isaacman~(1981) surveyed the central $2\\arcdeg \\times 4\\arcdeg$ ($\\ell \\times b$) of the GC in a search for planetary nebulae. He finds an \\emph{excess} number of sources as compared to that expected from extragalactic source counts, an excess he attributes to \\ion{H}{2} regions and planetary nebulae. That he finds an excess at all is notable, though, since our analysis predicts that angular diameters of extragalactic sources seen through the GC scattering region will be at least 1\\farcm5--3\\arcmin. The resolution of his survey was $0\\farcm4 \\times 2\\arcmin$, and he was able to detect sources with angular scales as large as 14\\arcmin. Thus, we attribute his excess to the fact that, where his survey overlapped the GC scattering region, it was desensitized by the intense GC scattering to a considerably lesser degree than our survey. Anantharamaiah et al.~(1991) used their observations at~0.327~GHz and a 0.408~GHz $\\log N$-$\\log S$ relation to conclude that the number of observed extragalactic sources within a 4~deg${}^2$ area centered on \\sgra\\ is consistent with the number expected from high-latitude source counts. Outside of the inner 1~deg${}^2$, our source counts are also consistent with the expected number of extragalactic sources. Although scattering of other sources, such as B1739$-$298 and B1741$-$312 is heavy, the predicted diameter of these sources is less than 10\\arcsec\\ at~0.327~GHz, comparable to the size of their beam, so that these sources would not have been resolved out. Gray et al.~(1993) surveyed the Sgr~E region ($\\ell = 358.7\\arcdeg, b = 0\\arcdeg$) at~0.843, 1.45, and~4.86~GHz. At 1.4~GHz the number of sources they find is consistent with that expected from the $\\log N$-$\\log S$ distribution. Figure~\\ref{fig:sclike} shows that the likelihood function for our field 358.9$+$0.5 does not favor a large amount of scattering, i.e., the number of sources in this field is consistent with that expected. Further, two of the sources observed in our VLBI experiment (\\cite{lc97}), B1739$-$298 and 1LC~358.439$-$0.211, are in this field. The former is heavily scattered, though not at a level sufficient for it to be seen through the \\sgra\\ scattering screen. We conclude that our source count results are in good agreement with previous source counts toward the \\hbox{GC}. The only exception occurs over the 1~deg${}^2$ region centered on \\sgra. This region has not been considered previously or has been subsumed into a much larger area. \\subsection{Physical Conditions in the Scattering Region}\\label{sec:gc.physical} Our global likelihood, Fig.~\\ref{fig:global} and \\S\\ref{sec:gc.global}, attained a maximum for the following parameter values: $\\delgc = 150$~pc, $0.5\\arcdeg \\le \\psi_\\ell \\lesssim 1\\arcdeg$, and $l_0^{2/3}\\te^{-1/2} = 10^{-7}$. In this section we consider whether a medium exists within the GC for which such parameter values are plausible. There is a wealth of observational data available for the \\hbox{GC}. We shall summarize those conclusions relevant to our study here; interested readers are referred to a number of recent reviews---Genzel, Hollenbach, \\& Townes~(1994); Morris \\& Serabyn~(1996); and Gredel~(1996)---and references within. Our criteria for the host medium of the density fluctuations are that the medium must have a sufficient density and that it must be capable of sustaining density fluctuations of the requisite magnitude. We establish our first criterion by estimating $\\nbar$ from the scattering diameters of GC sources using equations~(\\ref{eqn:measures}) and~(\\ref{eqn:smweight}). The diameters of \\sgra\\ and the OH masers require a weighted scattering measure of $S \\approx 10^2$~\\smu, equation~(\\ref{eqn:smweight}). Eliminating SM between equations~(\\ref{eqn:measures}) and~(\\ref{eqn:smweight}) and solving for $\\nbar$ yields \\begin{equation} \\nbar \\sim 10^3\\,\\mathrm{cm}^{-3}\\frac{1}{\\varepsilon\\sqrt{f}}\\left(\\frac{l_0}{1\\,\\mathrm{pc}}\\right)^{1/3}\\left(\\frac{\\delgc}{150\\,\\mathrm{pc}}\\right)^{-3/2}. \\label{eqn:delne} \\end{equation} Two factors could alter this estimate by about an order of magnitude. First, it is likely that $l_0 \\ll 1$~pc, which would \\emph{reduce} our estimate of $\\nbar$. Second, as we noted in \\S\\ref{sec:gc.global}, the similarity between the values of $\\delgc$ and $\\psi_\\ell\\dgc$ suggests that the density fluctuations fill the region and $f \\approx 1$. However, we might also associate $l_0$ with the characteristic size of a scattering cloudlet within the region. If the region contains few such cloudlets and $\\delgc/l_0 \\gg 1$, then $f \\ll 1$, and our estimate above would be a considerable \\emph{underestimate}. Yusef-Zadeh et al.~(1994) estimated that a typical line of sight might intersect only 10 or so scattering cloudlets. In any event, we conclude that the scattering medium must be dense, $n_{\\mathrm{e}} \\gtrsim 10^2$~cm${}^{-3}$. For comparison, Spangler~(1991) concludes that $n_{\\mathrm{e}} \\sim 1$~cm${}^{-3}$ for scattering regions in the Galactic disk. Our second criterion for the host medium is that it must be able to support density fluctuations of the required magnitude. This constraint has been lucidly reviewed by Spangler~(1991): The density fluctuations are presumed to arise from plasma turbulence. As this turbulence dissipates, it cannot heat the host medium at a rate that exceeds the medium's cooling capacity. This constraint is particularly acute in the situation we are proposing as the dissipation mechanisms considered by Spangler~(1991) scale as $l_0^{-a}$ with $a \\approx 1$. Since we are considering $l_0 < 1$~pc, the heating rates could be excessive. The dominant damping mechanisms for $l_0 < 1$~pc are linear Landau damping, ion-neutral collisions, and a parametric decay instability. The first two mechanisms scale as $l_0^{-2/3}$ while the latter scales as $l_0^{-1}$. In addition to their dependence on $l_0$, the damping rates depend on the large scale magnetic field, $\\Gamma \\propto B^2$; the Alfv\\'en wave speed, $\\va$; and the amplitude of the magnetic fluctuations, $\\Gamma \\propto (\\delta B/B)^2$ for linear Landau damping and ion-neutral collisions while $\\Gamma \\propto (\\delta B/B)^3$ for the parametric decay instability. Linear Landau damping also depends upon the angle of propagation with respect to the direction of $\\mathbf{B}$, $\\chi$, and the plasma $\\beta$. For values of these quantities appropriate for scattering regions in the Galactic disk, these damping mechanisms produce volumetric heating rates of $\\Gamma \\sim 10^{-23.5}$--$10^{-21.5}$~erg~s${}^{-1}$~cm${}^{-3}$. As Spangler~(1991) discussed, there are also a number of simplifications and additional assumptions which enter the calculation of these heating rates. Inferred magnetic field strengths in the GC are $B \\sim 1$~mG, or $10^3$ that of the field strength in the disk. To estimate $\\delta B/B$, we use (Cordes, Clegg, \\& Simonetti~1990) \\begin{equation} \\frac{\\delne}{n_{\\mathrm{e}}} \\sim \\left(\\frac{\\delta B}{B}\\right)^c \\label{eqn:bB} \\end{equation} with $c = 1$ for linear processes and $c = 2$ for non-linear processes like the parametric decay instability. Our first criterion for the host medium is that $\\delne/n_{\\mathrm{e}} \\le 1$. For definiteness, and to provide the largest possible value of the heating, we take $\\delta B/B \\sim 1$. Finally, although $B$ is much larger in the GC than in the Galactic disk, $n$ is also larger. As a result $\\va$ is larger than in the disk, but probably by no more than an order of magnitude. Thus, we expect $\\Gamma$ in the GC to be about a factor of $10^7$ larger that in the Galactic disk. The heating rate from linear Landau damping in scattering regions in the Galactic disk (\\cite{s91}) assumes the density fluctuations arise from obliquely propagating magnetosonic waves ($\\chi \\approx 6\\arcdeg$). More aligned propagation results in less damping. It is not clear if the GC environment would favor highly aligned propagation or not. The large values of $B$ in the GC are inferred, in part, from the system of non-thermal filaments and threads seen throughout the \\hbox{GC}. With only one exception, these filaments have no kinks or bends in them, even though they are observed to be interacting with molecular clouds having typical velocities of 10--100~km~s${}^{-1}$ (\\cite{ms96}). This rigidity could be an indication that only highly aligned propagation is allowed. If $\\chi$ is highly concentrated near 0\\arcdeg, then the heating from linear Landau damping would be unimportant and the heating rates could be two orders of magnitude lower than those quoted above. Alternately, as Spangler~(1991) noted, the distribution of $\\chi$ could be isotropic, but waves with large $\\chi$ would then damp quickly and the heating rate will be unchanged or even larger than what we assume. The presence of small-scale ($\\approx 0.1$~pc) magnetoionic cloudlets in the GC has already been inferred to explain large changes in the Faraday rotation measure of certain features (G~359.1$-$00.2, the ``Snake,'' \\cite{gnec95}; G~359.54$+$0.18, the non-thermal filaments, Yusef-Zadeh, Wardle, \\& Parastaran~1997), though the inferred density in these cloudlets, 0.3--10~cm${}^{-3}$, is less than our nominal estimate. In the Galactic disk a small body of observational evidence suggests that the magnetoionic medium responsible for Faraday rotation is also responsible for scattering and pulsar dispersion (\\cite{sc86}; Lazio, Spangler, \\& Cordes~1990; \\cite{ahmsrs96}), and the same may be true in the \\hbox{GC}. We now consider two models for the host medium. In both models, the scattering arises in thin layers on the surfaces of molecular clouds. Even if the filling factor, $f$, of these layers is not large, the \\emph{covering factor}, i.e., the probability that a line of sight through the GC will intersect one of these layers, can still be close to unity. \\subsubsection{Photoionized Surfaces of Molecular Clouds}\\label{sec:gc.skins} Over the region $|\\ell| \\lesssim 1\\fdg5$ and $|b| \\lesssim 0\\fdg5$, $n_{\\mathrm{e}} \\sim 10$~cm${}^{-3}$, as determined from single-dish recombination line and total intensity measurements (Matthews, Davies, \\& Pedlar~1973; \\cite{mp79}). Embedded within this large-scale region are smaller regions of much higher electron densities, $n_{\\mathrm{e}} \\sim 10^3$--$10^5$~cm${}^{-3}$, primarily within Sgr~A West and Sgr~B2 (e.g., \\cite{mpgy-z93}). Some of these high density regions are the photoionized surfaces (size $\\sim 10^{-4}$~pc) of molecular clouds ($n \\gtrsim 10^4$~cm${}^{-3}$) irradiated by the ambient radiation field (effective temperature $\\approx 35\\,000$~K). Yusef-Zadeh et al.~(1994) identified these molecular skins as the source of the scattering and associated their thicknesses with the outer scale, $l_0$; Gray et al.~(1995) suggested that the magnetoionic medium responsible for the Faraday rotation toward G~359.1$-$00.2 also results from these molecular clouds. This model suffers from at least three potential difficulties. First, the molecular skins are photoionized by a radiation field having a temperature of $T_{\\mathrm{eff}} \\sim 10^4$~\\hbox{K}. Our constraint on $l_0^{2/3}\\te^{-1/2}$ therefore requires $l_0 \\sim 10^{-7.1}$~pc. This value is considerably smaller than that derived by Yusef-Zadeh et al.~(1994) for the ionized molecular skins. However, in deriving their value for $l_0$, Yusef-Zadeh et al.~(1994) used a value of the ionizing flux appropriate to the inner few parsecs. The stellar density decreases as $r^{-2}$, so outside the inner few parsecs, the ionizing flux should be lower than that assumed by Yusef-Zadeh et al.~(1994). A lower ionizing flux would result in a smaller skin depth and bring their estimate and our estimate of $l_0$ into better agreement. A second potential difficulty with this model is that the medium would only barely be capable of cooling itself. If the outer scale is $l_0 \\sim 10^{-7}$~pc, then the heating rate from the damping of the plasma turbulence is $\\Gamma \\sim 10^{-13}$--$10^{-12}$~erg~cm${}^{-3}$~s${}^{-1}$. The cooling capacity of gas near $\\te \\sim 10^4$~K depends sensitively upon the fractional ionization and temperature (\\cite{dm72}). We estimate that a density of $n_{\\mathrm{e}} \\gtrsim 10^5$~cm${}^{-3}$ is required for these skins to be able to cool sufficiently in order to support the density fluctuations. This density is at the upper end of the range $10^3$--$10^5$~cm${}^{-3}$ inferred for the small-scale \\ion{H}{2} regions. In determining the cooling function of the medium, we have used results that assume a solar abundance. The metallicity in the GC could be as much as twice solar, leading to an increased cooling efficiency. The third difficulty is that the required value for $l_0$ in this model, $l_0 \\sim 3 \\times 10^{11}$~cm, is considerably smaller than that in the Galactic disk. In the Galactic disk, a stringent lower limit on the outer scale is $10^{13}$~cm, and it may be as large as $10^{18}$~cm (\\cite{ars95}). Although the physics for the generation and maintenance of small-scale density fluctuations is not well understood, we regard it as potentially troublesome that this model predicts such a small $l_0$. \\subsubsection{``Warm'' Interfaces}\\label{sec:gc.interface} X-ray observations have revealed a central, diffuse X-ray source with a (FWHM) size of $1\\fdg8 \\times 0\\fdg9$ (\\cite{ykkkt90}). Frail et al.~(1994) suggested that this X-ray emitting gas may be responsible for the scattering. Yusef-Zadeh et al.~(1997) suggested that this X-ray emitting gas is also the magnetoionic medium responsible for the Faraday rotation toward G~359.54$+$0.18. The density and temperature of this region are estimated at 0.05~cm${}^{-3}$ and $10^7$--$10^8$~\\hbox{K}. This region cannot itself be the host of the density fluctuations because of its low density, cf.~equation~(\\ref{eqn:delne}). However, this gas appears spatially coincident with the central zone of intense molecular emission and presumably abuts cooler gas in the clouds. We modify Frail et al.'s~(1994) proposal by identifying the interfaces where the GC molecular clouds are exposed to this ambient hot medium as the source of the scattering. We term these interfaces ``warm'' by analogy with McKee \\& Ostriker's~(1977) model for the \\hbox{ISM}. In that model cold clouds immersed in a hot ($10^6$~K) medium have $10^4$~K interfaces. In the GC densities and temperatures are 1--2 orders of magnitude higher, but we expect that clouds will still develop intermediate temperature interfaces. This model suffers from two of the same difficulties as the previous model. The X-ray emitting gas and molecular clouds appear to be in rough pressure equilibrium with $P \\sim 5 \\times 10^6$~K~cm${}^{-3}$ (\\cite{bs91}). Even though there are supersonic motions \\emph{within} the clouds, the clouds themselves ($v \\sim 10$--100~km~s${}^{-1}$) are moving subsonically with respect to the hot medium ($c_{\\mathrm{sound}} \\sim 1000$~km~s${}^{-1}$). Taking pressure balance to extend throughout the interface region, we find a density $n_{\\mathrm{e}} \\sim 5$--50~cm${}^{-3}$ for $\\te \\sim 10^5$--$10^6$~\\hbox{K}. From our likelihood results, the temperature within these interfaces implies an outer scale of $l_0 \\sim 10^{-6.5}$--$10^{-6}$~pc, which, in turn, implies an rms density, equation~(\\ref{eqn:delne}), of $\\delne \\sim 10$~cm${}^{-3}$. However, the cooling capacity of this medium is only $10^{-20}$~erg~cm${}^{-3}$~s${}^{-1}$. The predicted heating rate is $\\Gamma \\sim 10^{-13}$~erg~cm${}^{-3}$~s${}^{-1}$. The outer scale in this model remains troublesomely small. If the size of the interface region is set by thermal conduction, the portion of the interface with $\\te < 10^6~K$ has a size $\\lesssim 10^{-1}$~pc (\\cite{mc77}). Clearly if not all of the interface contributes to the scattering better agreement would be obtained between $l_0$ and the interface size. Still, the outer scale remains an order of magnitude smaller than its lower limit in the Galactic disk. One point in favor of this model is the distribution of the X-ray emitting gas as compared to that of the molecular clouds. The size of the X-ray emitting region is similar to the extent of the scattering region, approximately 1\\arcdeg. In contrast, the molecular cloud distribution extends over the range $-1\\arcdeg \\lesssim \\ell \\lesssim 2\\arcdeg$. If the scattering traced massive stars within these molecular clouds, as the photoionized molecular cloud skins model suggests, the scattering should extend further in longitude than it does. In this respect, the lack of enhanced scattering for \\hoh\\ masers in Sgr~B is particularly problematic. We stress the importance, and probable uniqueness, of the high density in the \\hbox{GC}. In the Galactic disk density fluctuations cannot be supported in media with $\\te \\sim 10^6$~K, a position with both theoretical and limited observational support (\\cite{s91}; \\cite{pc92}). Similarly, recent VLBI observations of 5 pulsars show no evidence for an enhanced level of turbulence at the boundary of the Local Bubble (\\cite{cr87}; Britton, Gwinn, \\& Ojeda~1996), potentially a local analog of an interface between a hot and cooler medium. However, the Local Bubble and ambient medium have densities a factor of $10^2$--$10^3$ smaller than that in the \\hbox{GC}. In summary, we use our likelihood results, \\S\\ref{sec:gc.global}, to constrain host media for the scattering material. Potential media include the photoionized skins of molecular clouds or the interface regions between the clouds and the ambient X-ray emitting gas. There are difficulties with both models: Both models overpredict the outer scale and appear to have some trouble supporting the required level of density fluctuations. Although we have been unable to make an unambiguous identification of the scattering medium with either medium, we favor the interface model, in part, because it shows a better correspondence between the spatial distribution of scattering and proposed host medium. \\subsection{Modification of the Taylor-Cordes Model}\\label{sec:tcmodel} The TC93 model modelled the global distribution of free electrons in the Galaxy with four components: an extended component, an inner Galaxy component, a component confined to the spiral arms, and a component local to the Gum Nebula. We now extend the TC93 model to include a GC component\\footnote{ As in TC93, the coordinate system has the $x$-axis directed parallel to $\\ell = 90\\arcdeg$, the $y$-axis toward $\\ell = 180\\arcdeg$, the $z$-axis toward $b = 90\\arcdeg$, and $R = \\sqrt{x^2 + y^2}$ is the Galactocentric radius.} (cf.\\ eqn.~[11] of TC93): \\begin{eqnarray} n_{\\mathrm{e}}(x,y,z) & = & n_1(R,z) + n_2(R, z) + \\sum_{j=1}^4 n_{\\mathrm{arm},j}(x,y,z) + n_{\\mathrm{Gum}}(x,y,z) \\nonumber \\\\ & + & n_{\\mathrm{GC}}g_{\\mathrm{GC}}(R)h_{\\mathrm{GC}}(z). \\end{eqnarray} The first four components are discussed at length in TC93. We focus on only the last component, that toward the \\hbox{GC}. Based on the estimate in equation~(\\ref{eqn:delne}) and our estimates for $l_0$, we take $n_{\\mathrm{GC}} = 10$~cm${}^{-3}$. Our estimate for $\\delgc$ is $\\delgc = 150$~pc. Heretofore, we have been treating the scattering region as a screen with sharp boundaries. It is more likely that the region has soft edges. We therefore adopt a radial dependence of \\begin{equation} g_{\\mathrm{GC}}(R) = \\exp\\left[-(R/0.150\\,\\mathrm{kpc})^2\\right]. \\end{equation} The latitude (or $z$) dependence of the screen is less well constrained. For definiteness we take \\begin{equation} h_{\\mathrm{GC}}(z) = \\exp\\left[-(z/0.075\\,\\mathrm{kpc})^2\\right] \\end{equation} corresponding to $\\psi_b = 0\\fdg5$. The resulting axial ratio for the electron density distribution is 0.5; the axial ratio for the X-ray distribution is also 0.5 and that of the molecular cloud distribution is 0.3 (\\cite{ms96}). In the TC93 model the relationship between the free electron density and the scattering measure produced by a line of sight of length $ds$ through those electrons is $d\\mathrm{SM} \\propto F n_{\\mathrm{e}}^2ds$. The parameter~$F$ is \\begin{equation} F = \\frac{\\zeta\\epsilon^2}{f}\\left(\\frac{l_0}{1\\,\\mathrm{pc}}\\right)^2, \\label{eqn:F} \\end{equation} where $\\zeta$ is the normalized variance of electron density fluctuations between cloudlets and their surroundings and $\\epsilon$ and~$f$ are as in equation~(\\ref{eqn:measures}). Taking $\\zeta \\sim \\epsilon \\sim 1$, our estimates for $l_0$ imply $F \\gtrsim 10^4$. Both of the models we have considered here have $f < 1$. For definiteness, we take $f \\sim 0.1$, recognizing that this may be an upper limit on~$f$. We therefore conclude $F \\gtrsim 10^5$. For comparison, the parameter~$F$ has a value of~0.4 in the solar neighborhood, 6 in spiral arms, and 40 in the Galaxy's inner few kiloparsecs. A value of $F \\sim 10^5$ produces an SM comparable to that suggested by the scattering diameters of \\sgra\\ and the OH masers. Their scattering diameters require a line-of-sight weighted scattering measure of $S \\approx 10^2$~kpc~m${}^{-20/3}$. Our results suggest $\\delgc \\approx 150$~pc; correcting for the line-of-sight weighting, equation~(\\ref{eqn:smweight}), implies that the GC has a scattering measure of $\\mathrm{SM} \\sim 10^{5.5}$~kpc~m${}^{-20/3}$. Integrating the TC93 expression for $d\\mathrm{SM}$ through the GC, with $F \\sim 10^5$, we find $\\mathrm{SM} \\sim 10^6$~kpc~m${}^{-20/3}$. Since the GC component is so localized, only for lines of sight through the GC are the results of TC93 altered. In this direction, however, the TC93 model underpredicts various quantities by a large amount. In the TC93 model, GC pulsars have $\\mathrm{DM} \\approx 600$--800~pc~cm${}^{-3}$; we predict that the DM will be somewhat larger, $\\mathrm{DM} \\approx 2000$~pc~cm${}^{-3}$, with approximately 1500~pc~cm${}^{-3}$ of that arising from the GC component itself. For comparison, the largest DM known is for PSR~B1758$-$23 with $\\mathrm{DM} = 1074$~pc~cm${}^{-3}$ (Manchester, D'Amico, \\& Tuohy~1985; \\cite{klmjds93}). Further, from Cordes \\& Lazio~(1997) the temporal broadening of pulses from pulsars seen through this region will be $350\\,\\nu_{\\mathrm{GHz}}^{-4}$~\\emph{seconds}, requiring high-frequency ($\\nu \\approx 10$~GHz) periodicity searches to detect pulsations. Although the DM we predict for GC pulsars is substantial, the dispersion smearing across a 1~GHz bandpass at 10~GHz ($\\approx 5$~ms) is comparable to the pulse broadening, so that only a small number of filterbank channels, e.g., 16, would be necessary to combat the dispersion smearing. \\subsection{Conclusions} We use a likelihood analysis to determine the following parameters of the GC scattering region: The GC-scattering region separation, $\\delgc$; the angular extent of the region, $\\psi_\\ell$ and $\\psi_b$; the outer scale on which density fluctuations occur, $l_0$; and the gas temperature, $\\te$. \\begin{itemize} \\item From the literature we have assembled a list of all sources toward the GC for which angular broadening has been measured. A subset of these sources is OH/IR stars, for which the spatial distribution about the GC is known. We construct a likelihood function for the angular broadening of OH/IR stars, utilizing this distribution, \\S\\ref{sec:broaden}. Masers within approximately 1\\arcdeg\\ of \\sgra\\ have diameters consistent with $\\delgc \\approx 150$~pc (Fig.~\\ref{fig:aggregate_as}). \\item The likelihood analysis of our source counts, \\S\\ref{sec:counts}, indicates that a deficit of sources occurs within approximately 1\\arcdeg\\ of \\sgra\\ and is caused by a scattering region within 500~pc of \\sgra\\ (Fig.~\\ref{fig:aggregate_sc}). The resulting scattering diameter, at least 20\\arcsec\\ at~1~GHz, causes extragalactic sources to be so broad as to be resolved out by our observations. \\item \\hoh\\ masers in and an extragalactic source near Sgr~B and an extragalactic source near Sgr~C do not show the extreme scattering of sources closer to \\sgra, indicating that the scattering region does not extend to more than 1\\arcdeg\\ in longitude. The latitude extent of the scattering region is poorly constrained, but is no more than 1\\arcdeg. \\item From the literature we have estimated the free-free emission and absorption toward five heavily scattered masers near \\sgra. The likelihood function is dominated by the free-free emission. The relevant parameters, $\\delgc$, $l_0$, and $\\te$, are not independent for this likelihood and only the product $\\delgc^{-2}l_0^{2/3}\\te^{-1/2}$ can be constrained (Fig.~\\ref{fig:aggregate_ff}). \\end{itemize} The global likelihood, formed by multiplying the individual likelihoods, is shown in Fig.~\\ref{fig:global}. The maximum likelihood estimates of the parameters are $\\delgc = 150$~pc, $0.5\\arcdeg \\le \\psi_\\ell \\lesssim 1\\arcdeg$, and $l_0^{2/3}\\te^{-1/2} = 10^{-7}$ with $l_0$ in pc and $\\te$ in \\hbox{K}. The parameter $\\psi_b$ was not well constrained and we adopt $\\psi_b = 0\\fdg5$. The close correspondence between $\\delgc$ and $\\psi_\\ell\\dgc$ suggests that the scattering region encloses the \\hbox{GC}. The GC scattering region produces a 1~GHz scattering diameter of 90\\arcsec, if the region is a single screen, or 180\\arcsec, if the region wraps around the GC, as appears probable. We modify the Taylor-Cordes model for the Galactic distribution of free electrons in order to include an explicit GC component. We predict that pulsars seen through this region will have a dispersion measure of approximately $2000$~pc~cm${}^{-3}$, of which approximately 1500~pc~cm${}^{-3}$ arises from the GC component itself, and suffer pulse broadening of $350\\,\\nu_{\\mathrm{GHz}}^{-4}$~\\emph{seconds}; pulsations will be detected only for frequencies above 10~GHz (\\cite{cl97}). As host media for the scattering we consider the photoionized surface layers of molecular clouds and the interfaces between molecular clouds and the $10^7$~K ambient gas. We identify the host medium by requiring that it be sufficiently dense to support density fluctuations of the required magnitude. We are unable to make an unambiguous determination, but we favor the interface model which predicts that the scattering medium is hot ($\\te \\sim 10^6$~K) and dense ($n_{\\mathrm{e}} \\sim 10$~cm${}^{-3}$). The X-ray interface model also shows better spatial agreement, when compared to the photoionized skin model, with the region over which the scattering is observed. This model is summarized graphically in Fig.~\\ref{fig:gc.summary}. The GC scattering region is likely to be unique in the Galaxy, probably because it is a high-pressure environment and can sustain densities and temperatures much higher than in the Galactic disk." }, "9804/astro-ph9804294_arXiv.txt": { "abstract": "{ Lithium is one of the few primordially produced elements. The value of the primordial Li is taken to be that observed in metal--poor dwarfs, where it is not contaminated by stellar Li sources which act on longer time scales. The atmospheric abundance is currently derived from the LiI $\\lambda\\lambda 6707 \\AA~$ resonance transition and the validity of the models employed has been questioned \\markcite{k95} (Kurucz 1995). In this letter we report the first detection of the Li I $\\lambda\\lambda 6104 \\AA~ 2^2P - 3^2D$ subordinate transition in the prototype population II star HD~140283. The same Li abundance of (Li/H) $=1.4\\times 10^{-10}$ is found consistent with both the resonance and subordinate lines. The two lines form at different depths in the atmosphere implying that the 1-D homogeneous atmospheric models used in the abundance determination are essentially correct. When coupled with the standard big bang yields, the Li in the halo dwarfs provides two solutions for the baryon-to-photon ratio $\\eta_{10}= n_{b}/n_{\\gamma} \\times 10^{10}$ and for the present baryon density $\\Omega_b h_{70}^2=0.0748\\eta_{10}$: a) a first solution at $\\eta_{10}\\approx 1.8$, consistent with the $\\eta_{10}$ implied by the high deuterium values $D/H\\approx 2\\times 10^{-4}$ observed in some quasar absorption systems \\markcite{webb} (Webb et al 1997) and b) a second solution at $\\eta_{10}$ $\\approx$ 4 which is consistent, within the errors, with the low deuterium D/H =$3.4\\times 10^{-5}$ measured in other quasar absorption systems\\markcite{burles} (Burles \\& Tytler 1998). } ", "introduction": "Lithium, together with D and $\\rm ^{3,4}He$, is one of the few elements produced by nuclear reactions in the first minutes after big bang\\markcite{wfh} (Wagoner, Fowler \\& Hoyle 1967). The observations of these elements and their extrapolation to the primordial values are consistent with the predictions of the standard primordial nucleosynthesis providing, together with the relic radiation and the expansion of the Universe, a robust support to the big bang theory. Recently, additional support to the primordial nature of Li in halo dwarfs has come from the observations of Li in metal-poor stars of the thick disc\\markcite{mbp97} (Molaro, Bonifacio \\& Pasquini 1997). This population is chemically and kinematically distinct from the halo, but has the same Li abundance of the halo. Minniti et al (1997) \\markcite{min97} claimed detection of Li, at the plateau level, in a metal-rich, but old star, belonging to the Galactic Bulge. Finally Li at the plateau level has also been detected in a star which was possibly born in an external galaxy and then accreted by the Milky Way\\markcite{mol97} (Molaro 1997). So far the Li abundance has been always obtained only from the analysis of the Li I $\\lambda\\lambda$ 6707 \\AA~ resonance doublet. This is not a very comfortable situation in the light of the importance of the determination of lithium abundances in stars for primordial nucleosynthesis, stellar structure and chemical evolution. Our ability to determine the Li abundance using simple plane-parallel homogeneous atmospheres, has been recently debated\\markcite{k95,kis97,gp97} (Kurucz 1995; Kiselman 1997; Gadun \\& Pavlenko 1997). The analysis of several lines, which sample different depths in the stellar atmosphere is crucial to test the correctness of the modelling. The one dimensional, homogeneous, static models which are currently employed may arise concern because they ignore the fine structure and hydrodynamic phenomena such as granulation which are seen on the Sun. The Li I $\\lambda\\lambda$ 6707 \\AA ~resonance transition is the only one readily available to spectroscopic observation. The strongest subordinate line at 6104 \\AA ~ is much fainter and blended with Fe I line and has been so far detected only in young T Tauri stars (Hartigan et al 1989)\\markcite{hart} and Li-rich giants (Merchant 1967, Wallerstein \\& Sneden 1982) \\markcite{merchant,ws}, were Li is more than about 1 dex more abundant owing to the Galactic Li production. \\par In this letter we report the detection of the Li I $\\lambda\\lambda$ 6104 \\AA ~transition in the spectrum of the metal--poor star HD140283. Both this line and the resonance line are consistent with the computations made using a one dimensional, homogeneous model atmosphere, thus increasing our confidence that this model represents a satisfactory average of the complex fine structure expected in metal--poor stars. The use of Li observed in halo dwarfs as an indicator of primordial abundance rests on the absence of any Li depletion. Depletion is predicted by non-standard models which take into account rotational mixing \\markcite{pin92} (Pinsonneault, Deliyannis \\& Demarque 1992) or diffusion \\markcite{vau95} (Vauclair \\& Charbonnel 1995), but these models predict a downturn of the hot side of the Li plateau and considerable dispersion. It seems that neither the downturn nor the large dispersion is present in the observations, which suggests that diffusion or rotational mixing do not affect significantly the Li observed at the stellar surface of metal--poor dwarfs. However, the downturn can be very small ($\\approx 0.2$ dex) for the purely diffusive case and a suitable choice of the mixing length parameter ($\\alpha=1.5$, see Deliyannis et al 1990\\markcite{deli90}) and the issue of intrinsic dispersion remains rather controversial with some positive claims. Ryan et al (1996)\\markcite{ryan} identify a triplet of stars (G064-012, G064-037, CD -33$^\\circ$ 1173) with similar colors, but different Li abundances by a factor of 2.5. Then there is the case of star BD+23 3912 which has a [Fe/H]$\\approx -1.3$ to $-1.5$ and a Li abundance which is about 0.20-0.36 dex higher than the plateau (Rebolo et al 1988, King et al 1996\\markcite{reb88,king96}). Moreover Boesgaard et al (1998)\\markcite{boesgaard} find differences of up to $\\approx 0.5$ dex among seven subgiants of M92 but the same objects show other chemical peculiarities, namely [Mg/Fe] is 0.55 dex lower and [Na/Fe] is 0.76 dex larger than in HD140283 (King et al 1998) \\markcite{king98} . ", "conclusions": "The LiI line forming regions lie in the upper part of the atmospheric convective zone where Li is mostly ionized due to its low ionization potential. This is why the determination of precise Li abundances requires accurate observations, accurate stellar effective temperature and an appropriate modeling of the atmosphere of a metal poor star. The model-atmospheres employed are one dimensional (1-D), with plane parallel geometry and ignore any inhomogeneity effect, such as granulation. Qualitative computations, based on a two-stream model atmosphere, suggested that the abundance of Li in halo dwarfs could be underestimated by as much as a factor of 10\\markcite{k95} (Kurucz 1995), but more recent calculations based on 2-D\\markcite{gp97} (Gadun \\& Pavlenko 1997) and 3-D\\markcite{kis97} (Kiselman 1997) atmospheric models show that effects of granulation on the LiI lines are much less important. Granulation effects in the atmosphere have a depth dependence and this should produce different effects in the resonance and subordinate doublets. As can be seen from figures 1 and 2, the same Li abundance reproduces satisfactorily both the 6104 \\AA ~and the 6707 \\AA ~doublets. The two transitions form at different depths in the stellar atmosphere: unit optical depth at wavelength 6707.761 \\AA~ is attained at $log (\\tau_{Ross})\\approx -0.57$, corresponding in our model to a local temperature of 5235 K, while at wavelength 6103.649 \\AA~ it is already attained at $log (\\tau_{Ross})\\approx -0.09$, or T=5915 K. The resonance line receives contributions from a more extended region than the subordinate line. Unit optical depth at the wavelength at which the residual intensity is 0.999, is attained at $log (\\tau_{Ross})\\approx -0.11$ for the resonance line, but at $log (\\tau_{Ross})\\approx -0.08$ for the subordinate line. Thus the subordinate line samples deeper and hotter layers than the resonance line, as shown in Fig. 3. \\par The lower level of the Li 6104 \\AA~ transition is the upper level of the 6707 \\AA $2^2S-2^2P$~ transition. Our synthetic spectra are computed under the LTE assumption and the consistency between the two lines implies a correct computation of the populations of the 2S, 2P and 3D levels. This is in agreement with the theoretical estimations which predict relatively small corrections for NLTE effects in the LI 6707 \\AA ~line \\markcite{carl94,pavmag} (Carlsson et al 1994; Pavlenko \\& Magazz\\`u 1996). Thus the detection of a subordinate LiI line, and its consistency with the abundance derived from the resonance 6707 \\AA~ doublet, provides support to the correctness of this Li abundance. The consistency of the abundances based on the LiI 6707 \\AA~ and 6104 \\AA~ transitions observed in HD 140283 supports the Li abundances measured in the population II stars, using 1-D model atmospheres, in the last decades. The new generation of large telescopes will allow to measure the Li 6104 \\AA ~Li I subordinate doublet in other much fainter population II stars, thus permitting to verify this consistency on the grounds of a statistically significant sample, and ultimately achieve a more accurate measurement of the primordial Li abundance. \\par Among the light elements produced in the first minutes after the big bang, Li is the only one which shows a non monotonic behaviour with $\\eta_{10}$, the so-called {\\sl Li-valley}, which reflects the different nuclear reactions which synthesize Li at different baryonic densities. The most recent measurement of the Li primordial abundance is $\\rm (Li/H)=1.73 \\pm 0.05_{stat} \\pm 0.2_{sys} \\times 10^{-10}$\\markcite{bm} (Bonifacio \\& Molaro 1997), which is the mean value of 41 halo stars for which precise effective temperatures, determined by means of the infrared flux method \\markcite{alonso} ({Alonso}, {Arribas}, \\& {Martinez-Roger} 1996), were available. The systematic errors, which dominate the error budget, come from a possible offset of $\\pm$ 75 K in the zero point of the temperature scale of cool stars. This Li abundance intercepts the primordial yields for two different values of $\\eta_{10}$, which unfortunately do not help in resolving the deuterium and helium controversies. Each solution for $\\eta_{10}$ obtained from Li is consistent with either the high-deuterium/low-helium\\markcite{webb,olive} (Webb et al 1997; Olive, Steigman \\& Skillman 1997) or the low-deuterium/high-helium \\markcite{burles,izo97} (Burles \\& Tytler 1998; Izotov, Thuan \\& Lipovetsky 1997). The lower $\\eta_{10}$ requires considerable D destruction to match the presently observed abundance in the local interstellar medium of the Galaxy. The higher $\\eta_{10}$ value is also consistent with the low deuterium (D/H $=3.9 (\\pm 1)\\times 10^{-5}$) derived from the 92 cm hyperfine transition emission towards the unprocessed Galactic anti-center \\markcite{chen97} (Chengalur, Braun \\& Butler Burton 1997)." }, "9804/astro-ph9804249_arXiv.txt": { "abstract": "Modifications in Friedmann-Lema\\^itre-Robertson-Walker (FLRW) Hubble diagrams caused by mass density inhomogeneities are used to illustrate possible effects on a determination of the mass parameter $\\OM$ and the cosmological constant $\\Lambda$. The values of these parameters inferred from a given set of observations depend on the fractional amount of matter in inhomogeneities and can differ from those obtained by using standard FLRW Hubble diagrams by as much as a factor of two. ", "introduction": "The pressure-free FLRW models of GR have, for a very long time, been assumed to adequately describe the large scale geometry of our universe but only recently have observational techniques become available which promise to determine values for all three of their requisite parameters $\\{H_0,\\ \\OM$, and $\\OL\\equiv {\\Lambda c^2/(3H_0^2)}\\}$ and implicitly test this assumption. As an example, corrected magnitudes and redshifts ($m$-$z$) for Type Ia supernovae (SNe Ia) are measured, plotted, and compared with theoretical $m(H_0,\\ \\OM,\\OL; z)$ curves computed for the FLRW models \\cite{PS1,PS2,GP}. Because these models are isotropic and homogeneous, and our universe appears quite inhomogeneous, modifications in the FLRW predictions have long been proposed and estimated \\cite{YZ,KR}. When a wide angle measurement of the CMB is made the average mass density of the universe (the FLRW value for $\\rho_0$) likely exists within the radio beam collected by the antenna. However, when small objects such as SNe Ia are observed at $z< 1$, a mass density significantly less than the average is `likely' to be in the observing optical path. In particular, if the underlying mass density approximately follows luminous matter (\\ie associated with bounded galaxies) then effects of a diminished mass density in the observing beams on relations like $m(\\OM,\\OL;z)$ are important. The majority of currently observed SNe Ia are not being seen through foreground galaxies and whether or not this is due to selection (rather than statistics) is not important. If the objects observed do not have the average FLRW mass density $\\rho_0$ in their foregrounds then the FLRW \\mz\\ relation does not apply to them. Ultimately some SNe Ia should exist behind foreground galaxies and for these, \\mz\\ should be computed using the lensing formulas. These formulas \\cite{BR,CJ} contain source-observer, deflector-observer, and source-deflector distances, respectively $D_s, D_d$, and $D_{ds}$, all of which depend on the mass density in the observing beam, {\\bf excluding} the deflector. These distances will not be given by the standard FLRW result if the observing beam contains less than the average FLRW mass density, but instead given by the `intergalactic' distance discussed here. ", "conclusions": "Numerous arguments have been made since the '60s against the existence of any effect on apparent magnitudes such as given here (see \\cite{KR2} for current refs. to dissenting opinions). These are almost all weak-lensing arguments, valid as long as density perturbations don't produce multiple images (or absorb photons), \\ie the FLRW result will coincide with the theoretical mean of weak lensing observations. Even if weak lensing arguments can be extended to strong lensing perturbations, \\eg by not resolving separate images, the theoretical mean of a distribution of magnitudes may be of little use in determining $\\OM$ or $\\OL$. If mass is as inhomogeneous as luminosity, the distribution of magnitudes is expected to be so skewed as to make the mean statistically insignificant. The `most likely' value should be far more useful in a determination of the cosmic parameters. However, to determine the most likely \\mz not only requires knowledge of the average mass density $\\rho_0$, it requires the modeling of galaxies masses etc. What can be determined by fixing a single additional parameter $\\nu$ (which proportions total FLRW mass density into an intergalactic component and a galactic component) is the `intergalactic' Hubble curve. This \\mz can be much more useful in determining $\\OM$ or $\\OL$ than the mean Hubble curve, \\eg if galaxies are not larger than they appear optically, the distribution of magnitudes (at a given redshift) is expected to peak much closer to the intergalactic value than to the mean value (see the numerical work \\cite{HD}). Additionally the effects of galaxy lensing can be controlled by simple selection. All that is required is that the observed SNe Ia are separated into those with foreground galaxies and those without. Those without should be fit to the intergalactic Hubble curve and those with (when any are found) could be included by correcting for lensing and/or by averaging in with the others to see if the weak-lensing FLRW value can be obtained. Only with enough unbiased and absorption corrected data will the FLRW Hubble curve be useful. The intergalactic Hubble curve contains the additional parameter $\\nu$; however, if luminous matter is essentially the whole story most matter is in galaxies and one can put $\\nu=2$ as a good approximation. Use of the intergalactic Hubble curve is then, in principle, no more involved than use of the standard FLRW Hubble curve. The division of $\\rho_0$ into galactic and intergalactic parts for gravitational-optics purposes seems simplistic but it is certainly less simplistic than ignoring optical effects of inhomogeneities altogether as is done by using FLRW." }, "9804/astro-ph9804280_arXiv.txt": { "abstract": "The history of the transition from a neutral intergalactic medium to one that is almost fully ionized can reveal the character of cosmological ionizing sources. In this talk I will discuss the implications for rival reionization scenarios of the rapid decline observed in the space density of quasars and star-forming galaxies at redshifts $z\\gta 3$. The hydrogen component in a highly inhomogeneous universe is completely reionized when the number of ionizing photons emitted in one recombination time equals the mean number of hydrogen atoms. At $z\\sim 5$, the local character of the UV metagalactic flux allows one to define a {\\it critical} emission rate of hydrogen-ionizing photons per unit comoving volume, ${\\dot{\\cal N}}_{\\rm ion}=10^{51.5\\pm 0.3}\\ndotunits$. Models based on photoionization by bright QSOs and/or young galaxies with star formation rates in excess of $0.3-1\\sfr$ appear to fail to provide the required number of hydrogen-ionizing photons at these redshifts by large factors. If stellar sources are responsible for keeping the universe ionized at $z\\approx 5$, the rate of star formation per unit comoving volume at this epoch must be comparable or greater than observed at $z\\approx 3$. ", "introduction": "The existence of a filamentary, low-density intergalactic medium (IGM) which contains the bulk of the hydrogen and helium in the universe is predicted as a product of primordial nucleosynthesis and of hierarchical models of gravitational instability with ``cold dark matter'' (CDM) (Cen \\etal 1994; Zhang \\etal 1995; Hernquist \\etal 1996). The application of the Gunn-Peterson constraint on the amount of smoothly distributed neutral material along the line of sight to distant objects requires the hydrogen component of the diffuse IGM to have been highly ionized by $z\\approx 5$ (Schneider \\etal 1991), and the helium component by $z\\approx 2.5$ (Davidsen \\etal 1996). It thus appears that substantial sources of ultraviolet photons were present at $z\\gta 5$, perhaps low-luminosity quasars or a first generation of stars in virialized dark matter halos with $T_{\\rm vir}\\gta 10^4\\,$K (Couchman \\& Rees 1986; Ostriker \\& Gnedin 1996; Haiman \\& Loeb 1997; Miralda-Escud\\`e \\& Rees 1997). Early star formation provides a possible explanation for the widespread existence of heavy elements in the IGM (Cowie \\etal 1995), while reionization by QSOs may produce a detectable signal in the radio extragalactic background at meter wavelengths (Madau \\etal 1997). Establishing the character of cosmological ionizing sources is an efficient way to constrain competing models for structure formation in the universe, and to study the collapse and cooling of small mass objects at early epochs. While the nature, spectrum, and intensity of the background UV flux which is responsible for maintaining the intergalactic gas and the Ly$\\alpha$ clouds in a highly ionized state at $z\\lta 3$ has been the subject of much debate in the last decade, it is only in the past few years that new observations have provided reliable information on the presence and physical properties of the sources and sinks (due to continuum opacities) of UV radiation in the interval $3\\lta z\\lta 5$. In this talk I will focus on the candidate sources of photoionization at early times and on the time-dependent reionization problem, i.e. on the history of the transition from a neutral IGM to one that is almost fully ionized. The starting point of this study can be found perhaps in the simple realization that the {\\it breakthrough epoch} (when all radiation sources can see each other in the Lyman continuum) occurs much later in the universe than the {\\it overlap epoch} (when individual ionized zones become simply connected and every point in space is exposed to ionizing radiation), and that at high redshifts the ionization equilibrium is actually determined by the {\\it instantaneous} UV production rate. In the following I will adopt an Einstein-de Sitter universe ($q_0=0.5$) with $H_0= 50h_{50}\\,\\kmsmpc$. ", "conclusions": "" }, "9804/astro-ph9804233_arXiv.txt": { "abstract": "\\noindent Nonlinear time-dependent calculations are being carried out in order to study the evolution of vertically-integrated models of non-selfgravitating, transonic accretion discs around black holes. In this paper we present results from a new calculation for a high-$\\alpha$ model similar to one studied previously by Honma, Matsumoto \\& Kato who found evidence for limit-cycle behaviour connected with thermal instability. Our results are in substantial agreement with theirs but, in our calculation, the disc material does not always remain completely optically thick and we include a suitable treatment for this. We followed the evolution for several cycles and determined the period of the cycle as being about 780 seconds. Advective cooling is dominant in the region just behind the outward-moving peak of surface density. The behaviour of this model is significantly different from what we saw earlier for low-$\\alpha$ models (which we discussed in a previous paper) and we contrast and compare the two situations. ", "introduction": " ", "conclusions": "" }, "9804/astro-ph9804005_arXiv.txt": { "abstract": "We have obtained intermediate resolution spectra of eleven candidate brown dwarf members of the Pleiades open cluster using the Keck II telescope and LRIS spectrograph. Our primary goal was to determine the location of the ``lithium depletion edge\" in the Pleiades and hence to derive a precise age for the cluster. All but one of our 11 program objects have radial velocities appropriate for Pleiades members, have moderately strong \\Ha\\ emission, and have spectral types M6 to M8.5 as expected from their (R-I)$_c$\\ colors. We have constructed a color-magnitude diagram for the faint end of the Pleiades main sequence, including only stars for which high S/N spectra in the region of the lithium $\\lambda$6708$\\AA$\\ absorption line have been obtained. These data allow us to accurately determine the Pleiades single-star lithium depletion edge at I$_{c0}$\\ = 17.80, (R-I)$_{c0}$\\ = 2.20, spectral type = M6.5. By reference to theoretical evolutionary models, this converts fairly directly into an age for the Pleiades of $\\tau$\\ = 125 Myr. This is significantly older than the age that is normally quoted, but does agree with some other recent estimates. ", "introduction": "The Pleiades was recognized in the 1980's as the best open cluster to attempt to identify brown dwarfs (Stauffer \\etal 1989; Jameson \\& Skillen 1989) because of its fortuitous combination of proximity, youth and richness. A number of brown dwarf candidates were identified in those and subsequent papers, usually on the basis of photometry obtained from deep imaging surveys in two or more colors. A means to establish that a candidate brown dwarf was at least near, if not necessarily below, the hydrogen burning mass limit was proposed by Rebolo, Mart\\'{\\i}n and Magazz\\`u (1992; hereafter RMM). Below about 0.065 \\MSUN, brown dwarfs should never develop core temperatures sufficient for lithium ignition, and thus for objects less massive than this we should find a lithium abundance the same as for the interstellar medium from which the object formed, independent of the object's age. For slightly higher masses, lithium acts as an age scale because the length of time it takes for the core to reach 2.5x10$^6$\\ K is a sensitive function of mass. Because these stars are fully convective, once the core temperature exceeds the necessary limit, the entire lithium content of the star should be exhausted rapidly and thus be reflected in an observable change in the photospheric lithium abundance. The current generation of theoretical models make specific predictions about the time evolution of this lithium depletion boundary. For example, D'Antona \\& Mazzitelli (1997) predict that at ages 30, 70 and 140 Myr, the lithium depletion edge should occur at 0.17 \\MSUN, 0.09 \\MSUN\\ and 0.07 \\MSUN, respectively. Other recent models by Baraffe \\etal (1998) and Burrows \\etal (1997) make nearly identical predictions of the variation of this lithium depletion boundary with age. Indeed, Bildsten \\etal (1997) and others have argued that the age for an open cluster derived in this manner should be better than by any other method. The initial attempts to detect lithium in very low mass open cluster members were not successful (RMM; Marcy, Basri \\& Graham 1994; hereafter MBG94). However, Basri, Marcy \\& Graham (1996; hereafter BMG96) eventually detected lithium in PPL 15, a Pleiades brown dwarf candidate originally identified by Stauffer, Hamilton \\& Probst (1994). Rebolo et al. (1996) later showed that two other Pleiades members about 1 magnitude fainter than PPL 15 also have strong lithium absorption. BMG96 derived an age for the Pleiades based on the presence of lithium in PPL 15 but the absence of lithium in another only slightly brighter Pleiades member (HHJ3). However, Basri \\& Mart\\'{\\i}n (1998) have recently provided evidence that suggests that PPL 15 is a nearly equal-mass spectroscopic binary and thus its individual components would be $\\sim$0.75 mag fainter. This allows the location of the lithium boundary to be considerably fainter than had been assumed. The published data therefore no longer constrain the age of the Pleiades nearly as well as one would like because too few stars have been measured spectroscopically in the magnitude range of interest. In this paper, we report on the results of a program to determine lithium abundances for a number of candidate Pleiades members with apparent magnitudes chosen to bracket the possible magnitude range within which the lithium depletion boundary might be located. ", "conclusions": "The goal of this project was to precisely determine the location of the lithium depletion edge in the Pleiades, and hence to determine an accurate age for the cluster. Figure 2 shows a color-magnitude diagram for very low mass Pleiades members for which lithium data are available. The lithium data are from Oppenheimer \\etal (1997 = Opp97), MBG94, BMG96, Rebolo \\etal (1996 = Reb96) and from this paper. The (R-I)$_c$\\ colors for the Opp97 stars are from Stauffer \\etal (1995), where we have converted these spectroscopic (V-I)$_c$\\ colors to (R-I)$_c$\\ via a relation derived from the Gliese catalog M dwarfs in Leggett (1992). The (R-I)$_c$\\ colors for some stars in Reb96 and BMG96 are from PC2 indices provided by Mart\\'{\\i}n \\etal (1996) and our calibration of PC2 vs. (R-I)$_c$. Two points for CFHT-PL-15 are shown: (R-I)$_c$\\ = 2.24 as derived from its PC2 index, and (R-I)$_c$ = 2.41 as derived from its VO index. Using either color, CFHT-PL-15 lies below the main sequence defined by the other stars, possibly indicating that it is a non-member. Because it has detected lithium and a radial velocity compatible with Pleiades membership, we prefer to believe it is a member and that the inferred color is anomalous (either intrinsically or due to measurement error). Our primary conclusions are unaffected by how we interpret this star. Finally, the dashed line in Figure 2 is a field star ZAMS derived from the M dwarf photometry provided by Leggett (1992). The location of the stars in Figure 2 show a well-defined correlation with the measured lithium equivalent widths. Bluer than (R-I)$_c$\\ = 2.2, none of the stars have detected lithium while redder than that color all of the measured stars have lithium. Similarly, fainter than I$_{c0}$\\ = 17.8 all the stars have detected lithium, whereas brighter than that limit all but one of the stars have no lithium. Finally, there is a trend for lithium equivalent width to increase going to lower inferred mass (i.e. fainter and redder). We further believe that the dispersion in I magnitude at a given color is to a large extent real and is primarily an indication that some of the observed stars are photometric binaries. In particular, we suggest that HHJ6 (I$_{c0}$\\ = 16.93, R-I$_c$\\ = 2.18), Lick-PL1, PPL1, and CFHT-PL-12 are good candidates to be photometric binaries. CFHT-PL-12 also has considerably stronger \\Ha\\ emission than the other Pleiades stars observed - possibly indicating it is a short-period binary or that we have observed it during a flare. The attribution of PPL1 as a nearly equal mass binary would explain why it has strong lithium absorption despite an I$_c$\\ magnitude equal to or brighter than two other cluster members with no detected lithium. An alternative explanation would be that there is a significant age spread in the Pleiades and these four over-luminous stars are the youngest in our sample. By comparison to theoretical models (e.g. D'Antona \\& Mazzitelli 1997), their displacement about 0.5 mag above other stars of the same color would require them to be more than 50 Myr younger than the other stars, which we believe is unlikely (see, for example, discussions of this issue by Soderblom \\etal 1993 and Steele \\& Jameson 1995). Based on the above interpretation of Figure 2, we determine that the single star lithium depletion edge in the Pleiades is at I$_{c0}$\\ = 17.8 $\\pm$ 0.1 or (R-I)$_{c0}$\\ = 2.20 $\\pm$ 0.05. The uncertainty estimates are not rigorous, and arise mostly from the uncertainties in the absolute calibration of the photometry in Bouvier \\etal (1998). For theoretical evolutionary models which incorporate realistic model atmospheres as their outer boundary condition, and hence which can predict observational colors and magnitudes for brown dwarfs, it is possible to convert directly the empirical lithium depletion edge to an age estimate for the Pleaides. In Figure 3, we plot the absolute I$_c$\\ magnitude of the lithium depletion boundary (defined here as the point where lithium has been depleted by a factor of 100) for the most recent models of Baraffe \\etal (1998) as a function of age. To place our empirically measured point into this diagram, we assume that the cluster has (m-M)$_o$\\ = 5.60 (r $\\sim$ 130 pc), and A$_I$\\ = 0.06 (c.f. Pinsonneault \\etal 1998), leading to M(I$_c$) = 12.2 $\\pm$\\ 0.15 for the lithium depletion boundary, where we have assumed plausible but again unrigorous 1$\\sigma$\\ uncertainties of 0.1 mag for the distance modulus and 0.03 mag for the extinction and that the uncertainties add in quadrature. The age derived in this way is then 125 $\\pm$\\ 8 Myr, where the uncertainty estimate only comes from propagating the 0.15 mag uncertainty of the boundary through the model shown in Figure 3. In order to try to assess the model dependence of this estimate, we have also made a similar calculation for theoretical evolutionary models by Burrows \\etal (1997) and D'Antona \\& Mazzitelli (1997). Those models do not provide R and I magnitudes, so instead we have used the (R-I)$_c$ color to estimate an I$_c$-band bolometric correction (we adopted the Monet \\etal (1992) BC$_I$ vs. (V-I)$_c$\\ relation, and a conversion from (V-I)$_c$\\ to (R-I)$_c$\\ based on data in Leggett 1992). In that manner, we estimate that the lithium depletion boundary in the Pleiades is at M(Bol) = 11.99. Comparing this number to the predictions of the two theoretical models, we get an age for the Pleiades of 130 Myr based on the D'Antona \\& Mazzitelli (1997) calculations and 125 Myr for the Burrows \\etal (1997) models. The most commonly quoted age for the Pleiades is of order 70-80 Myr (Patenaude 1978; Mermilliod 1981). However, models with a relatively large amount of convective core overshoot can yield much larger ages, as was originally shown by the models of the Padova group (c.f. Mazzei \\& Pigatto 1989, who derived an age for the Pleiades of 150 Myr). Other recent models give ages intermediate between these values (e.g. 100 Myr for Meynet, Mermilliod \\& Maeder 1993 and $\\geq$\\ 120 Myr for Ventura \\etal 1998). Given the disagreement over the age derived from the upper main sequence turn-off, it is particularly useful to have an independent means to derive the age from low mass stars. Using the lithium detection in PPL15 and the non-detection of lithium in HHJ3, BMG96 estimated the age of the Pleiades to be 115 Myr. Our new result pushes this age even slightly older, but more importantly does so using many more stars and thus provides a much better defined age from the lithium data. Finally, we note that we have chosen to use the ``traditional\" distance scale to the Pleiades (r $\\sim$\\ 130 pc). That distance conflicts with the new Hipparcos distance to the Pleiades of about 116 pc (van Leeuwen \\& Ruiz 1997; Mermilliod \\etal 1997). We have done this because we believe that the Hipparcos distance for the Pleiades is not correct, as has been discussed in Pinsonneault \\etal (1998) and Soderblom \\etal (1998). The zeroth order effect of simply using the Hipparcos distance and keeping everything else the same would lead to an even older Pleiades age of about 140 Myr." }, "9804/astro-ph9804097_arXiv.txt": { "abstract": "Multiwavelength observations of high energy flare in 1996 from 3C 279 seems to favour the so called mirror model between different inverse Compton scattering models proposed as a possible explanation of gamma-ray emission in blazars. We performed kinematic analysis of the relativistic blob - mirror system and found that only part of the mirror located very close to the jet axis (very likely inside the jet cone) can re-emit soft photons which serve as a target for production of $\\gamma$-rays by relativistic electrons in the blob. Since the presence of well localized scattering mirror inside the jet is problematic, this makes problems for the mirror model. The time scale and the shape of the $\\gamma$-ray flare should reflect, in terms of the mirror model, the blob dimensions and the longitudinal distribution of relativistic electrons inside the blob. For the $\\gamma$-ray light curve of the type observed in 1996 from 3C 279, i.e. the rising time of the flare during a few days with a sharp cut-off towards the end of the flare, the density of electrons inside the blob should increase exponentially starting from the front of the blob and reach maximum towards the end of the blob. Such distribution of electrons is difficult to explain in a model of a relativistic shock moving along the jet, which would rather inject electrons more efficiently at the front of the blob with a trail of particles on its downstream side. ", "introduction": "About 50 blazars have been detected by the Compton Gamma Ray Observatory in the MeV - GeV energy range (Fichtel et al. 1994, von Montigny et al. 1995, Thompson et al. 1995, Mukherjee et al. 1997), and 3 blazars, of the BL Lac type, are discovered in the TeV $\\gamma$-rays by the Whipple Observatory (Punch et al. 1992, Quinn et al. 1996, Catanese et al. 1997). These blazars can reach very high $\\gamma$-ray luminosities which are variable on time scales as short as a part of a day, in the case of optically violent variable quasars, or even several minutes, in the case of BL Lacs. These observations strongly suggest that $\\gamma$-ray emission from blazars is collimated towards the observer within a small angle as a result of relativistic motion of plasma in the jet or directional acceleration of particles. High energy processes occurring in blazars are popularly explained in terms of the inverse Compton scattering (ICS) model in which $\\gamma$-rays are produced in ICS of soft photons by electrons in a blob moving relativistically along the jet. Different modifications of this general model mainly concern the origin of soft photons, i.e. whether they come internally from the blob in the jet (synchrotron self-Compton (SSC) model, e.g. Maraschi et al. 1992, Bloom \\& Marscher 1993), directly from the disk (e.g. Dermer et al. 1992, Bednarek et al. 1996a,b), are produced in the disk but reprocessed by the matter surrounding the disk (external comptonization (EC) model, e.g. Sikora et al. 1994, Blandford \\& Levinson 1995), or produced in the jet but reprocessed by the matter surrounding the jet (the so-called mirror model, Ghisellini \\& Madau~1996, henceforth GM). In this last paper it is mentioned that SSC model and external comptonization of photons produced by the broad line region clouds (BLR) illuminated by the disk (EC model) may also contribute to the $\\gamma$-ray emission producing a first $\\gamma$-ray pre-flare. For the SSC model the amplitude of the $\\gamma$-ray variation is expected to be proportional to the square of the variation observed in IR-optical-UV energy range. For the EC model the $\\gamma$-ray emission should vary linearly with the low energy synchrotron emission. Such behaviour is not observed in the case of the 1996 flare from 3C 279 in which the $\\gamma$-ray variation is more than the square of the synchrotron variation. Moreover, in the $\\gamma$-ray light curve of this flare (see Fig.~1 in Wehrle et al. 1997), there is no clear evidence for a double peak structure which could eventually correspond to the first $\\gamma$-ray flare produced in terms of SSC or EC models and the second $\\gamma$-ray flare produced in terms of the mirror model. Therefore, although the SSC model can not be completely rule out, Wehrle et al. (1997) concludes that the mirror model is favourite by the multiwavelength observations of a strong flare in February 1996 from 3C 279 since it predicts $\\gamma$-ray flare with observed features. In this paper we test the mirror model by comparing predictions of the kinematic analysis with the observational results. The possible contributions from SSC and EC models to the $\\gamma$-ray production during this flare are neglected since, as we mentioned above, there is no observational support for their importance. Simultaneous analysis of all these models will require an introduction of additional free parameters (density of electrons in the blob, the perpendicular extend of the blob, definition of the disk radiation) which are not all well constrained by the observations. ", "conclusions": "We discuss details of the mirror model proposed by Ghisellini \\& Madau. This model seems to be favourite by the multiwavelength observations of the $\\gamma$-ray flare in 1996 from 3C 279 (Wehrle et al. 1997). Based on the analysis of the kinematics of the emission region (a blob moving relativistically along the jet) we come to the conclusion that only relatively small part of the mirror is able to re-emit soft photons which serve as a target for production of $\\gamma$-rays. For the parameters of the $\\gamma$-ray flare observed in 1996 from 3C 279, the radius of this part of the mirror should be comparable to the longitudinal extend of the blob. It has to be of the order of $2 \\times 10^{16}$ cm in order to be consistent with the rising time of the flare. This part of the mirror should lay inside the jet cone provided that its opening angle is of the order of $\\sim 1/\\gamma$. As mentioned in Ghisellini \\& Madau (GM), the physical processes in the jet may prevent the presence of the well localized mirror inside the jet. The calculations of density of photons re-emitted by the mirror are done by Ghisellini \\& Madau (see Fig.~2 in GM) in a time independent picture which do not take into account the dynamics of the blob. As a consequence they integrate over the parts of the mirror at distances from the jet axis which are much larger than the maximum distance $h_{\\rm u}$ (Eq.~(\\ref{eq17})), found in our dynamical (time dependent) analysis. The photon densities seen by the blob can not be directly compared with that ones obtained by us in a time dependent version of the mirror model. Ghisellini \\& Madau results are only correct for the continuous (time independent) flow of relativistic plasma along the jet axis but overestimates the density of soft photons seen by the relativistic electrons in the blob with limited longitudinal extend. The relativistic blobs in blazars has to be confined to the part of the jet in order to produce the $\\gamma$-ray flares with the observed rising time scale. We computed the $\\gamma$-ray light curves expected in the dynamical version of the mirror model for different distribution of relativistic electrons inside the blob and assuming that the density of electrons in the blob changes during propagation along the jet. Slowly rising $\\gamma$-ray flux with sudden cut-off towards the end of the flare, as observed in 3C 279, is obtained in the case of inhomogeneous blob with electron densities exponentially rising towards the end of the blob. Such electron distribution is difficult to understand in the popular scenario for $\\gamma$-ray production in which relativistic shock moves along the jet. It seems that such shock should rather inject relativistic electrons with high efficiencies close to the front of the blob, with the trail of electrons on its downstream side (Kirk, Rieger \\& Mastichiadis~1998). However the $\\gamma$-ray light curve expected in this case is different than observed during the flares in the blazar 3C 279. Since $\\gamma$-rays are produced in a region which is close to the mirror, therefore the shape of the $\\gamma$-ray light curve is not very sensitive on the variations of the density of electrons during the time of propagation of the blob between the base of the jet and the mirror. Of course the absolute $\\gamma$-ray fluxes produced by the blobs with different evolutions of electron densities in time may differ significantly. The $\\gamma$-ray light curves presented in Figs.~\\ref{fig2} show very sharp cut-offs towards the end of the flare due to our assumption on the negligible thickness of the mirror. In fact, the observed width of the peak in the $\\gamma$-ray light curve of 3C 279, of the order of $t_m\\sim 1$ day (see Fig.~1 in Wehrle et al. 1997), may be related to the time in which relativistic blob is moving though the mirror with the finite thickness. If this interpretation is correct then the thickness of the mirror has to be limited to $\\rho_{\\rm m}\\approx c t_{\\rm m} \\beta (1+\\beta)\\gamma^2 \\approx 4\\times 10^{17}$ cm which is comparable to the distance of the mirror from the base of the jet. In this analysis we do not consider production of $\\gamma$-rays in terms of the SSC and EC models simultaneously with the mirror model since there is no clear evidence of their importance in the $\\gamma$-ray light curve and the multiwavelength spectrum observed in 1996 from 3C 279 (Wehrle et al. 1997). The $\\gamma$-ray light curves reported in Figs.~2 show only relative change of the $\\gamma$-ray flux with time. They are not straightforwardly dependent on the parameters of the blob (the magnetic field, electron density, blob perpendicular extend, disk radiation) which are not uniquely constrained by the observations. The SSC and EC models will require to fix these parameters in order to guarantee reliable comparisons." }, "9804/astro-ph9804268_arXiv.txt": { "abstract": "Observations of galactic black hole candidates made by the instruments aboard the Compton GRO in the hard X-ray and $\\gamma$-ray bands have significantly enhanced our knowledge of the emission properties of these objects. Understanding these observations presents a formidable challenge to theoretical models of the accretion flow onto the compact object and of the physical mechanisms that generate high-energy radiation. Here we summarize the current state of observations and theoretical interpretation of the emission from black hole candidates above 20 keV. The all-sky monitoring capability of BATSE allows, for the first time, nearly continuous studies of the high-energy emission from more than a dozen black hole candidates. These long-term datasets are particularly well-suited to multiwavelength comparison studies, from the radio upward in frequency (Zhang et al. 1997a, these proceedings). Energy spectral evolution and/or spectral state transitions have been observed from many of the black hole candidates. Moderately deep searches of the galactic plane suggest a deficit of weak $\\gamma$-ray transients. Such population studies have implications for the origin of black hole binaries and the nature of accretion events. Observations above 50 keV from OSSE demonstrate that in the $\\gamma$-ray band there exist two spectral states that appear to be the extensions of the X-ray low (hard) and high (soft), or perhaps very high, states. The former state, the ``breaking'' state, cuts off with e-folding energy $\\sim$100 keV and has its peak luminosity near this energy; thus substantial corrections need to be made to historical estimates of the bolometric luminosity of black holes in the ``low'' state. In contrast, in the X-ray high (soft) state, the luminosity peaks in the soft X-rays and the spectrum extends with an unbroken power law, even up to energies above 500 keV in some cases. COMPTEL has detected emission above 750 keV from Cyg X-1 and the transient GRO~J0422+32. In both cases the data suggest that an additional weak, hard spectral component is required beyond that observed by OSSE at lower energies, although the precise spectral form is yet to be determined. The breaking $\\gamma$-ray spectrum can be well modeled by Comptonization of soft photons from the accretion disk in a hot thermal plasma. However, recent studies of the combined X-ray and $\\gamma$-ray spectrum of Cyg~X-1 and GX339--4 cast severe doubts on the simple geometry of a hot corona overlying a thermal accretion disk. Furthermore, timing studies of the former source are inconsistent with spectral formation by Compton scattering in a uniform, compact hot cloud, suggesting instead a decline in electron density with increasing radius. The power-law $\\gamma$-ray spectral state creates more significant theoretical challenges, particularly in explaining the lack of a break at energies exceeding the electron rest mass. It has been suggested that in the X-ray high (soft) state, the high-energy emission arises from bulk-motion Comptonization in the convergent accretion flow from the inner edge of the accretion disk. Such a process can conceivably generate the $\\gamma$ ray spectrum extending without a cutoff, if the accretion rate approaches that of Eddington. ", "introduction": "The most reliable evidence for the presence of a black hole in a binary system comes from determination of a mass function through optical measurements of the radial velocity of the companion star. If the resulting lower limit on the mass of the compact object exceeds 3$\\Msun$, the upper limit for the mass of a stable neutron star based on current theory, then one can reasonably assume that the compact object is a black hole. There are at least nine X-ray binary systems with minimum mass estimates exceeding 3$\\Msun$, of which three (Cyg~X-1, GRO~J0422+32, and GRO~J1655--40) have been clearly detected by GRO instruments. Other objects are identified as BHCs based on the similarity of their high-energy spectra and rapid time variability to those of Cyg~X-1. Such classification is, of course, somewhat tenuous. Before neutron stars and black holes can be reliably distinguished based on their X-ray and $\\gamma$-ray spectra, the full range of spectral forms from both classes must be observed and characterized. Extensive knowledge of the X-ray emission of these objects has accumulated in the literature, but the broad nature of the $\\gamma$-ray emission is only now coming to light, with the high sensitivity of current-generation instruments. The instruments of the Compton GRO have made extensive observations in the hard X-ray and $\\gamma$-ray bands of galactic black hole candidates (BHCs). With its all-sky capability, BATSE has monitored emission on a nearly continuous basis from at least three persistent sources (Cyg~X-1, 1E1740.7--2942, GRS~1758--258) and eight transients (GRO~J0422+32, GX339--4, N Mus 1991, GRS~1716--249, GRS~1009--45, 4U~1543--47, GRO~J1655--40, and GRS~1915+105). Lightcurves are presented below in Fig. \\ref{lightcurve}. OSSE has made higher-sensitivity, pointed observations of all of these sources, spectra of which appear below in Fig. \\ref{two_state}. COMPTEL has detected emission above 750 keV from Cyg~X-1 and GRO~J0422+32. To date, there have been no reported detections of galactic BHCs by EGRET. The French coded-aperture telescope Sigma on the Russian Granat spacecraft has imaged at least a dozen BHCs, including most of those in the list above, but with the addition of TrA~X-1, GRS~1730--312, and GRS~1739--278. The latter two objects were weak transients discovered during a multi-year survey of the galactic center region and have been classified as BHCs by their outburst lightcurves and the hardness of their spectra (Vargas et al. 1997). In this survey, Sigma regularly detected the persistent, variable sources 1E1740.7--2942 and GRS~1758--258, both of which are classified as BHCs on spectral grounds. The most striking result from Sigma observations of BHCs is the detection of broad spectral features below 500 keV from 1E1740.7--2942 (Sunyaev et al. 1991, Bouchet et al. 1991, Churazov et al. 1993, Cordier et al. 1993) and N Mus 1991 (Goldwurm et al. 1992, Sunyaev et al. 1992). These features have been interpreted as thermally broadened and red-shifted annihilation radiation from the vicinity of the compact object. ", "conclusions": "" }, "9804/astro-ph9804118_arXiv.txt": { "abstract": "We present results of \\ASCA\\ deep exposure observations of the hardest X-ray source discovered in the \\ASCA\\ Large Sky Survey (LSS) project, designated as AX~J131501$+$3141. We extract its accurate X-ray spectrum, taking account of the contamination from a nearby soft source (AX~J131502$+$3142), separated only by 1$'$. AX~J131501$+$3141 exhibits a large absorption of $N_{\\rm H} = (6^{+4}_{-2})\\times 10^{22}$ \\NHUNIT\\ with a photon index $\\Gamma = 1.5^{+0.7}_{-0.6}$. The 2--10 keV flux was about $5\\times 10^{-13}$ \\FLUXUNIT\\ and was time variable by a factor of 30\\% in 0.5 year. From the highly absorbed X-ray spectrum and the time variability, as well as the results of the optical follow-up observations (\\cite{Akiyama98}), we conclude that AX~J131501$+$3141 is a type 2 Seyfert galaxy. Discovery of such a low flux and highly absorbed X-ray source could have a significant impact on the origin of the cosmic X-ray background. ", "introduction": "Since the discovery of the Cosmic X-Ray Background (CXB) more than 30 years ago (\\cite{Giacconi62}), its origin has been a long standing puzzle. With the \\ROSAT, $\\sim$70\\% of the CXB below 2 keV has been resolved into point sources, more than 60\\% of which are type 1 active galactic nuclei (AGNs) (\\cite{Vikhlinin95}; \\cite{Hasinger96}; \\cite{McHardy98}; \\cite{Hasinger98}). Origin of the CXB above the 2 keV band, however, is less clear due to the absence of the imaging instrument in this energy band. One problem in the hard X-ray band, often referred as the spectral paradox (e.g. \\cite{Fabian92}), is that the X-ray spectrum of the CXB in the 2--10 keV band is harder than that of the typical type 1 AGN, which is presumably the main contributor to the CXB below 2 keV. The 2--10 keV X-ray spectra of type 1 Seyfert galaxies (most of the bright AGNs) are approximated by a power-law with a mean photon index of 1.7 (\\cite{Mushotzky93}), which is significantly steeper than that of the 2--10 keV CXB of about 1.4 (\\cite{Gendreau95}). This fact implies that the origin of the CXB above 2 keV differs, at least in part, from that below 2 keV. In addition, 20\\% of the total energy of the CXB is contained in the 2--10 keV band, whereas only a few percent of the CXB is contained below 2 keV (see review by Fabian \\& Barcons (1992), Hasinger (1996)). Thus, the 2--10 keV band would be an essential energy range to solve the origin of the CXB. \\ASCA\\ is the first satellite with the capability of the hard X-ray (up to 10keV) imaging and spectroscopy, hence is presently the best satellite to investigate the CXB in the 2--10 keV band. In the \\ASCA\\ Large Sky Survey project (LSS: \\cite{Inoue96}; \\cite{Ueda98a}), a continuous field of 7 deg$^2$ near the north galactic pole was surveyed with a sensitivity higher than any previous surveys in this energy band. Ueda \\etal\\ (1998a) resolved a significant fraction of the CXB, about 30\\% of the CXB, into discrete sources at a sensitivity limit of $F_{\\rm X}\\sim 10^{-13}$ \\FLUXUNIT\\ (2--10 keV). The mean photon index in the 2--10 keV band for these resolved sources ($F_{\\rm X}=$(1--4)$\\times 10^{-13}$ \\FLUXUNIT\\ in 2--10 keV) was found to be $\\Gamma = 1.5\\pm 0.2$. This result is consistent with the idea that the photon index approaches to that of the CXB, $\\Gamma\\sim 1.4$, as the source flux decreases. However, due to limited photon statistics, the spectral information of the resolved sources was too poor to address the nature of individual X-ray sources. The hardest source in the LSS (hereafter we refer it as the ``LSS hardest source'') was found to show a photon index of $\\Gamma\\sim -0.2$ with no correction of an absorption (\\cite{Ueda96}). However, it is unclear whether the apparent hard spectrum is due to a large absorption or due to flatness of the intrinsic spectrum. The LSS hardest source, which was found in an unbiased survey, would provide us a good opportunity to investigate the nature of faint and hard sources which could significantly contribute to the CXB above 2 keV. Hence, we have performed follow-up \\ASCA\\ and optical observations on the LSS hardest source. This paper reports results of the \\ASCA\\ deep exposure observations, while those of the optical observations are given by Akiyama \\etal\\ (1998). ", "conclusions": "We extracted the accurate spectrum of the LSS hardest source, AX~J131501$+$3141, taking account of the contamination from the nearby soft source, AX~J131502$+$3142, from which no significant X-ray emission was found in the LSS (Ueda 1996). We found that AX~J131501$+$3141 exhibits a large absorption of $N_H = (6^{+4}_{-2})\\times 10^{22}$ \\NHUNIT\\ with a photon index $\\Gamma = 1.5^{+0.7}_{-0.6}$. It showed a long-term time variability between two observations separated by 0.5 year. While the photon index of AX~J131501$+$3141 is consistent with the canonical value of type 1 AGNs (e.g., Mushotzky 1993), its absorption column density is larger than that of typical type 1 AGN by more than an order of magnitude, although a part of type 1 AGNs, about 10\\% of them (\\cite{Schartel97}), shows a column density larger than $5\\times 10^{22}$ \\NHUNIT. It is important that this source is selected fully unbiasedly. Hence, its X-ray properties should provide a key to understand the general nature of the missing hard X-ray populations which constitute the CXB above 2 keV. Two major possibilities have been proposed to account for the apparent hard spectrum of the CXB: one is to introduce large absorptions of sources (e.g., \\cite{Awaki91}), and the other is to consider populations of sources with intrinsically flat spectra (e.g., \\cite{Morisawa90}; \\cite{Matteo97}). Our results of the LSS hardest source strongly suggest that highly absorbed sources play an important role in considering the origin of the hard X-ray background. The large absorption of $6\\times 10^{22}$ \\NHUNIT, the photon index of $\\Gamma\\sim 1.5$, and the time variability are common properties seen in type 2 Seyfert galaxies. In fact, systematic studies of type 2 Seyfert galaxies by Awaki \\etal\\ (1991) and Ueno (1996) revealed that they commonly show large absorptions of $\\sim 10^{23}$ \\NHUNIT\\ and photon indices of 1.5--1.7. Akiyama \\etal\\ (1998) found one bright optical galaxy with $B=17.25$ mag near the center of the X-ray error circle of 0.5$'$ radius in the optical follow-up observations. No other optical source with the flux larger than $B=22.4$ mag is found in the error circle. Akiyama \\etal\\ (1998) performed spectroscopic observations of the bright galaxy and found that ratios of emission lines are similar to those found in type 2 Seyfert galaxies. The redshift of this galaxy was determined to be 0.07. From the redshift, the observed flux in the 2--10 keV band, $5\\times 10^{-13}$ \\FLUXUNIT, can be converted to the absorption corrected luminosity of $L_{\\rm X}\\sim 2\\times 10^{43}\\ {\\rm erg~s}^{-1}$. This luminosity is consistent with those of Seyfert galaxies. Thus, we identify AX~J131501$+$3141 found in the unbiased X-ray survey as a type 2 Seyfert galaxy. Using the LogN-LogS relation in Hasinger \\etal\\ (1998), we estimate that the chance coincidence between AX J131501$+$3141 and AX J131502$+$3142 is $\\sim$ 3\\%. However, these two sources have probably no physical correlation, because AX J131502$+$3142 is likely to be a QSO\\footnote{ In the optical imaging observations of R- and B-band by Akiyama \\etal\\ (1998), we found two point-like optical sources located at about 10$''$ north from the center of the error circle with 30$''$ radius for the soft source (AX J131502$+$3142). They are very close to each other and their magnitudes are comparable. Total magnitudes of the two sources are B=20.8 mag and R=19.6 mag; B-R color is 1.21. Since both the optical color and the optical to soft X-ray flux is consistent with that for type-1 AGNs, it is possible that one of them is a quasar which is also responsible for AX J131502$+$3142. However there are fainter optical sources (R$\\gtrsim$22) in the error circle and the optical identification for AX J131502$+$3142 is not clear yet. } which is more distant than the new type 2 Seyfert AX J131501$+$3141. Awaki (1991), Madau, Ghisellini \\& Fabian (1994) and Comastri \\etal\\ (1995) predicted that the combination of type 1 and type 2 AGNs can reproduce the CXB spectrum, based on the unified AGN scheme (e.g., \\cite{Antonucci93}). In the scheme, type 1 and type 2 AGNs are essentially the same objects, observed from different viewing angle. These type 2 AGNs, which exhibit apparently fainter and harder X-ray spectra than those of type 1 AGNs, should become detectable as the detector sensitivity increases. Although we have examined only one sample from the LSS at this moment, the result is encouraging not only for the unified AGN scheme, but also for solving the origin of the CXB." }, "9804/astro-ph9804162_arXiv.txt": { "abstract": "We present observations of NGC 4038/39 in the [\\ion{C}{2}] 158 \\micron \\/ fine structure line taken with the MPE/UCB Far-infrared Imaging Fabry-Perot Interferometer (FIFI) on the KAO. A fully sampled map of the galaxy pair (without the tidal tails) at 55\\arcsec \\ resolution has been obtained. The [\\ion{C}{2}] emission line is detected from the entire galaxy pair and peaks at the interaction zone. The total [\\ion{C}{2}] luminosity of the Antennae is $L_{\\rm [C II]} = 3.7 \\times 10^{8} L_{\\sun}$, which is about 1\\% of the far-infrared luminosity observed with IRAS. The main part of the [\\ion{C}{2}] emission probably arises from photodissociation regions (PDRs), and a minor fraction may be emitted from \\ion{H}{2} regions. A small part of the [\\ion{C}{2}] emission comes from standard cold neutral medium (CNM); however, for high temperature ($T \\sim 100$~K) and high density ($n_{\\rm H} \\sim 200$~cm$^{-3}$) about one third of the observed [\\ion{C}{2}] emission may originate from CNM. From PDR models we derive densities of the order of $\\sim 10^{5}$~cm$^{-3}$ and far-UV (FUV) intensities of $460\\chi_{\\circ}$, $500\\chi_{\\circ}$, and $240\\chi_{\\circ}$ for the PDRs in the interaction zone, NGC~4038, and NGC~4039, respectively. However, PDRs with densities of the order of $\\sim 10^{2}$~cm$^{-3}$ and FUV intensities of the order of $\\sim 100\\chi_{\\circ}$ could also explain the observed [\\ion{C}{2}] emission. The minimum masses in the [\\ion{C}{2}] emitting regions in the interaction zone and the nuclei are a few $\\times 10^{7}~M_{\\odot}$. A comparison with single dish CO observations of the Antennae shows a [\\ion{C}{2}] to CO intensity ratio at the interaction zone a factor of 2.6 lower than usually observed in starburst galaxies, but still a factor of about 1.3 to 1.4 higher than that at the nuclei of NGC 4038/39. Therefore, no global starburst is taking place in the Antennae. [\\ion{C}{2}] emission arising partly from confined starburst regions and partly from surrounding quiescent clouds could explain the observed [\\ion{C}{2}] radiation at the interaction zone and the nuclei, though the star formation activity toward the nuclei is lower. Accordingly there are small confined regions with high star formation activity in the interaction zone and with a lower star formation activity in the nuclei. This supports the high density and high FUV intensity for the PDRs in the interaction zone and the nuclei. ", "introduction": "The galaxy pair NGC 4038/39 (Arp 244) is an interacting system in an early stage of merging at a distance of about 21 Mpc from our own galaxy. On long-exposed images in the optical (e.g. Arp 1966) the interaction is clearly visible because of the tails (``Antennae'') emerging from two uniformly luminous, partly overlapping ovals and because of the dwarf galaxy that appears to have formed at the tip of the southern tail through the interaction (Zwicky 1956, Schweizer 1978, Mirabel, Dottori, \\& Lutz 1992). The tails contain about 70~\\% of the total amount of \\ion{H}{1} in the system (van der Hulst 1979). Short-exposed images (Laustsen, Madsen, \\& West 1987) reveal hints of the interaction on somewhat smaller scales; the distorted arrangement of H$\\alpha$ knots and the velocity distribution of the individual knots lead Rubin \\etal \\ (1970) to conclude there is an interaction of two rotating galaxies. Computer simulations carried out in the classical paper of Toomre \\& Toomre (1972) and later by Barnes (1988) can account for the present morphological appearance of the system quite well by assuming an interaction of two rotating spiral galaxies. Spectra of the nuclei of both galaxies taken in the optical range do not look like pure starburst spectra but consist of a composition of early-type stars and late giants (Keel \\etal \\ 1985). The detection of bright near-infrared peaks at the nuclei lead Bushouse \\& Werner (1990) to the same result. Their images of the Antennae in J- and R-band and in H$\\alpha$ show also the same pattern of bright knots in the surroundings of NGC 4038 and in the bridge connecting both galaxies. The Antennae system as a whole shows a relatively low star forming efficiency according to the ratio $L_{\\rm IR} /M({\\rm H}_2) \\approx 8.55$ measured by Young \\etal \\ (1986); it is only a factor of 3 higher than in the Milky Way. The ratio determined by Sanders \\& Mirabel (1985) is almost twice as high. However, they observed a smaller region in CO and therefore probably underestimated the molecular mass. Comparison of the Antennae with the sample of interacting and isolated galaxies of Young \\etal \\ (1986) shows NGC 4038/39 to have characteristics more like isolated galaxies. Measurements in the radio continuum at 1.5~GHz and 4.9~GHz of NGC~4038/39 (Hummel \\& van der Hulst 1986) reveal a number of discrete knots which coincide in general with H$\\alpha$ knots, and an underlying diffuse component. This diffuse component has a steep spectral index on average which indicates non-thermal emission, and the peak of the diffuse radio emission is at the dust patch near the overlapping region. The discrete radio knots account for roughly 35~\\% of the total radio emission and have a spectral index of $\\alpha \\approx -0.5$ on average, probably due to a thermal contribution (Hummel \\& van der Hulst 1986). Interferometric observations of the Antennae in CO ($1 \\to 0$) by Stanford \\etal \\ (1990) show three main concentrations of CO emission. Two are associated with the nuclei and the third with the interaction zone. The overlap region is the strongest CO source and contains $\\approx 10^{9}~M_{\\sun}$ of gas, roughly as much H$_{2}$ as both nuclei together. Based on 10$\\mu$m and H$\\alpha$ data, the authors have calculated a star forming rate of $5 M_{\\sun}$~yr$^{-1}$, and consequently the life time of the molecular gas is $2 \\times 10^{8}$~yr. Single dish observations in CO were made at the interaction zone and both nuclei of the Antennae by Aalto \\etal \\ (1995). The ratio of the emission lines of $^{12}$CO and $^{13}$CO measured in the nuclei and the overlapping region of the Antennae is similar to that found in the central regions of ``normal'' starburst galaxies (Aalto \\etal \\ 1995). An excellent tracer of star formation activity in galaxies is the strong [\\ion{C}{2}] 158~$\\mu$m $^{2}$P$_{3/2}\\to \\, ^{2}$P$_{1/2}$ fine structure line which arises mainly from photodissociation regions (PDRs) created by far-ultraviolet photons from hot young stars impinging on nearby dense interstellar clouds (Crawford \\etal \\ 1985, Stacey \\etal \\ 1991). In combination with CO and FIR observations, the [\\ion{C}{2}] emission can be used with PDR models (Tielens \\& Hollenbach 1985, Wolfire, Hollenbach, \\& Tielens 1989, Wolfire, Tielens, \\& Hollenbach 1990) to derive densities and far-UV intensities and estimates of the star formation activity. Extragalactic surveys of [\\ion{C}{2}] emission (Crawford \\etal \\ 1985, Stacey \\etal 1991) concentrated mainly on nuclei while more recent observations have imaged individual galaxies to study the distribution of [\\ion{C}{2}] emission on large scales. [\\ion{C}{2}] images of M83 (Geis \\etal \\ 1998) and NGC~6946 (Madden \\etal \\ 1993) demonstrate that the emission is extended at least over the full optical extent and often follows the distribution of the FIR and CO within the disk of the galaxies. In the case of NGC~6946 [\\ion{C}{2}] emission beyond the optical extent of the galaxy has been found. This [\\ion{C}{2}] emission has been attributed to diffuse gas and not to PDRs. Therefore we also estimate the possible contribution of [\\ion{C}{2}] emission from neutral atomic gas and from ionized gas in NGC~4038/39. Because of its relative proximity, NGC~4038/39 is a unique source for carrying out spatially resolved measurements of the [\\ion{C}{2}] line in an interacting system. We present the results of our imaging spectroscopy study of the [\\ion{C}{2}] line in NGC~4038/39 and compare them with observations obtained with ISO. ", "conclusions": "We present a map of NGC 4038/39 in the [\\ion{C}{2}] 158 \\micron \\ fine structure line. [\\ion{C}{2}] emission is detected over the optical extent of the system of galaxies and peaks at the interaction zone. The total luminosity of the [\\ion{C}{2}] line is $3.7 \\times 10^{8} L_{\\sun}$ which is about 1\\% of the FIR luminosity of the Antennae. Only a negligible fraction of the observed [\\ion{C}{2}] emission can originate in the WNM, if conditions are similar to Galactic atomic clouds. Under normal conditions the [\\ion{C}{2}] emission from standard CNM makes only a small contribution to the total [\\ion{C}{2}] emission, however it may rise to $\\slantfrac{1}{3}$ of the total [\\ion{C}{2}] emission for individual positions. Only minor fractions of the [\\ion{C}{2}] emission at the interaction zone and the nuclei can arise from \\ion{H}{2} regions. PDRs are the dominant source for the [\\ion{C}{2}] emission. We estimate minimum hydrogen masses associated with the [\\ion{C}{2}] emitting region of $1.9 \\times 10^{8} M_{\\sun}$ for the entire merging system and $6.8 \\times 10^{7} M_{\\sun}$, $3.2 \\times 10^{7} M_{\\sun}$, and $3.7 \\times 10^{7} M_{\\sun}$ within one beam centered at the interaction zone, NGC~4038, and NGC~4039, respectively. Assuming a single emission component in the beam we derive a density of the [\\ion{C}{2}] emitting gas in PDRs of $1 \\times 10^{5}$ cm$^{-3}$ for the interaction zone and of $2 \\times 10^{5}$ cm$^{-3}$ and $1 \\times 10^{5}$ cm$^{-3}$ for NGC~4038 and NGC 4039, respectively, and a far-UV intensity of $450 \\chi_{\\circ}$ for the interaction zone, $500 \\chi_{\\circ}$ for NGC~4038, and $250 \\chi_{\\circ}$ for NGC~4039. The derived beam filling factor of the emission from the PDRs is 20\\% for the interaction zone and 10--15\\% for the nuclei. However, the PDR model also allows a solution with low density ($\\sim 10^{2}$~cm$^{-3}$), high beam filling factor ($\\sim 50$\\%), and low FUV intensity ($\\sim 100 \\chi_{\\circ}$). The low, beam averaged [\\ion{C}{2}]/CO ratio of 2350 toward the interaction zone and the even lower ratios at the nuclei indicate that no global starburst is going on either in the area surrounding the interaction zone or in the nuclear region. The high-excitation lines observed with ISO SWS which trace the starburst must therefore arise from a small, confined region in the interaction zone. This result is also supported from the observations with ISOCAM. Therefore on the scale of our [\\ion{C}{2}] beam a single emission component for the PDR is only a crude approximation. Using interferometric and single dish CO observations and an expected [\\ion{C}{2}]/CO ratio for starburst regions and for quiescent clouds, we constructed a two-component model consisting of a confined starburst region and of molecular clouds enveloping the starburst. From this model we find that the [\\ion{C}{2}] emission at the interaction zone originates partly from confined starburst regions and partly from surrounding quiescent clouds. If we apply this model to the nuclei we get also enhanced star formation activity in NGC~4039 with a low [\\ion{C}{2}]/CO ratio for the quiescent clouds but only moderate star forming activity in NGC~4038. This two-component model supports the high-density solution for the PDRs. Future investigation of the Antennae in the [\\ion{N}{2}] 205 \\micron \\ fine structure line would be very helpful to further disentangle the origins of the [\\ion{C}{2}] line. Also observations at higher spatial resolution in the FIR regime (e.g. with FIRST and SOFIA) would be a great step forward to investigate this and other spatially very complex objects in more detail." }, "9804/astro-ph9804024_arXiv.txt": { "abstract": "Understanding the properties of interstellar turbulence is a great intellectual challenge and the urge to solve this problem is partially motivated by a necessity to explain the star formation mystery. This review deals with a recently suggested inversion technique as applied to atomic hydrogen. This technique allows to determine 3D turbulence statistics through the variations of 21~cm intensity. We claim that a radio interferometer is an ideal tool for such a study as its visibility function is directly related to the statistics of galactic HI. Next, we show how galactic rotation curve can be used to study the turbulence slice by slice and relate the statistics given in galactic coordinates and in the velocity space. The application of the technique to HI data reveals a shallow spectrum of the underlying HI density that is not compatible with a naive Kolmogorov picture. We show that the random density corresponding to the found spectrum tends to form low contrast filaments that are elongated towards the observer. ", "introduction": "The properties of the interstellar medium strongly suggest that it is turbulent. Here turbulence is understood as unpredictable temporal behavior of nonlinear systems as preached by J. Scalo (1985, 1987). The importance of turbulence in molecular clouds and its relation to star formation has long been appreciated (Dickman 1985). Recent progress in numerical simulations of molecular cloud dynamics (see Ostriker, this volume) indicates the intrinsic connection between the turbulence in different phases of the interstellar medium (McKee \\& Ostriker 1977). In what follows we shall mostly discuss the turbulence in atomic hydrogen (HI), although the formalism presented here is applicable to other spectral lines. Statistical description is a nearly indispensable strategy when dealing with turbulence and a big advantage of statistical techniques is that they extract underlying regularities of the flow and reject incidental details. Kolmogorov notion of energy cascade from large to small scales has been proved an extremely valuable concept and Kolmogorov spectrum of turbulence has been measured in various media. At the same time, astrophysical turbulence, unlike that in incompressible fluids, is a much more complicated phenomenon, and therefore one cannot {\\it a priori} hope that Kolmogorov's (1941) description is adequate (cf. Armstrong et al. 1995). Energy injection at small scales, shocks, compressibility may make interstellar turbulence spectrum much more informative, and we should expect to see deviations from the boring $-11/3$ slope. The advantage of using 21~cm emission data is that a continuum of separations between data points is available. This property is shared by diffuse emission in other spectral lines, but 21~cm measurements allow to disregard dust adsorption. As our review deals with HI studies within the galactic plane we do not discuss in detail interesting results obtained for HI in Large Magelanic Clouds (Spicker \\& Feitzinger 1988a,b). To avoid possible misunderstanding we should stress that Spicker \\& Feitzinger (1988a,b) deal with velocity fluctuations, while only intensity fluctuations are available when one studies HI turbulence in galactic plane. Statistics of random velocity and density fields may be different and therefore a direct comparison of the results obtained for these fields may be misleading. Being limited in space we refer the interested reader to the earlier reviews on interstellar turbulence, among which the one by Dickman (1985) can serve as an excellent introduction to the basic statistical techniques. A more advanced reader will enjoy a thoughtful analysis of problems associated with the statistical analysis of observational data given in Houlahan \\& Scalo (1990). Important aspects of the statistical analysis are discussed, for instance, by Dickman \\& Kleiner (1985), Roy \\& Joncas (1985) Perault et al. (1986), O'Dell \\& Casta\\~{n}eda (1987), Rickett (1988), Van Langevelde et al. (1992) Kitamura et al. (1993), Meisch \\& Bally (1994), Armstrong, Rickett, \\& Spangler (1995), Wallin, Watson, \\& Wyld (1998) and by the contributors to the present volume. A brief discussion of the very early studies of interstellar statistics can be found in Lazarian (1992). Studies of interstellar turbulence frequently deal with samples which are not statistically homogeneous (see Miesch \\& Bally 1994). Indeed, whenever individual molecular complexes are studied, the statistics (especially at large separations) may be dominated by regular gradients rather than the random component. To eliminate the inhomogeneous component, various types of spatial filtering are used (see Zurfleh 1967). These problems are alleviated for HI studies, since molecular clouds tend to be localized objects in sharp contrast to more pervasive distribution of atomic hydrogen. Further on we shall deal with the two point structure functions and power spectra. Naturally, one cannot place pickup devices at different points of HI. Instead, only the 21-cm intensity fluctuations with pointwise emissivity integrated along the lines of sight are available. It is obvious, that given a statistical description of the transparent emitting astrophysical medium, it is possible to predict statistical properties of the observable diffuse emission (Kaplan \\& Pikelner 1970), which would correspond to the solution of the {\\it forward problem}. However, more important is to solve the {\\it inverse problem}, namely, to deduce the 3D statistics of HI from observations. These issues are dealt with in sections 2 and 3. In section 4 we discuss the application of the technique to interferometric data. The galactic rotation curve allows one to study turbulence slice by slice. This slicing, however, is far from trivial (see section 4). Indeed, topologically disconnected blobs of HI can overlap in the velocity space if their velocities are the same. We show that interferometric study can potentially provide the information about both random density and velocity fields. Addressing the issue of HI topology we show that HI with the measured spectrum of density fluctuations forms low contrast filaments (section 5) and these filaments are elongated towards the observer due to the presence of velocity fluctuations. ", "conclusions": "The following are the principal conclusions of this review: \\begin{itemize} \\item The application of a newly developed inversion technique reveals a shallow spectrum of HI density in galactic plane. The spectrum is different from the Kolmogorov one, which may be indicative of non-trivial physics involved. \\item The spectrum of HI density is anisotropic in velocity space with velocity fluctuations altering the statistics along the line of sight. The degree of anisotropy and the spectrum in velocity space depend on the scale under study. The spectrum become isotropic for scales larger than 200-300~pc. \\item Interferometers are useful tools for studying HI turbulence in galactic plane and may provide the spectrum of 3D random density field and useful information on random velocity field. \\item Atomic hydrogen with Gaussian density and the shallow spectrum corresponding to observations forms low contrast filaments. These filaments are anisotropic in the velocity space and directed towards an observer. \\end{itemize}" }, "9804/astro-ph9804206_arXiv.txt": { "abstract": "As part of a continuing study of the effect of cluster environment on the star formation properties of galaxies, we have undertaken an H$\\alpha$ objective prism survey of the nearby cluster, Abell 1060. We detect 33 galaxies in emission, 24 of which are cluster members. We present new radial velocity measurements and H$\\alpha$+[\\ion{N}{ii}] equivalent widths and fluxes for a number of these galaxies. We distinguish between galaxies with diffuse and compact emission, the latter having been associated in previous work with a disturbed morphology of the galaxy and most likely resulting from tidally-induced star formation from galaxy--galaxy or cluster--galaxy interactions. The fraction of cluster spirals in Abell 1060 detected with compact emission agrees with the expected fraction for a cluster of its richness, as derived from results of a previous survey of 8 clusters. Some of the detected cluster early-type spirals exhibit anomalously high global H$\\alpha$ equivalent widths, as compared to galaxies of similar type in the field. ", "introduction": "\\label{resint} \\noindent The effect of cluster environment on the star formation properties of galaxies has long been a matter of debate. While some studies have suggested a lower star formation rate for cluster spirals as compared to the field (e.g. \\citeNP{gisler,dress85}), other work has suggested a similar or enhanced rate, particularly for early-type spirals(e.g. \\citeNP{kenn84,gav91,moss93,enacs}). With the discovery that in distant rich clusters there is a high fraction of blue, star-forming galaxies, often with unusual morphology suggestive of mergers and tidal interactions (e.g. \\citeNP{lavhen86,tom88,couch94}), there is renewed interest in tidally-induced star formation by mergers and interactions in nearby clusters. In order to address these issues, Moss, Whittle and co-authors have completed an objective-prism survey of eight nearby clusters of galaxies to detect global \\mbox{H$\\alpha$+[\\ion{N}{ii}]} emission as an indicator of the current rate of massive star formation. The survey technique is described by \\citeN{mwi}, hereafter MWI, and initial results have been discussed by \\citeN{moss93} (see also \\citeNP{moss90,moss95,moss96,moss97}). We have extended this survey to a ninth cluster, the Hydra I cluster, Abell 1060. Abell 1060 is one of the nearest of the Abell clusters, at a redshift of $z\\sim$0.01, and is the nearest large cluster beyond the Virgo and Fornax clusters. It has a high spiral fraction (e.g. \\citeNP{solanes92}) and is a relatively poor cluster, with a low density intracluster medium \\cite{lowmush} and low X-ray luminosity \\cite{edge91}. Since it is the nearest of the clusters surveyed by us so far, it can be surveyed to a fainter limit in absolute magnitude. However, its proximity means that it has a large projected diameter on the sky, with one Abell radius, \\mbox{$1\\, r_{A}=2\\fdg3=1.5\\, h^{-1}$ \\rm{Mpc}} (where $h$ is defined in terms of the Hubble constant $H_{0}=100h\\ \\rm{km\\ s}^{-1}\\rm{Mpc}^{-1}$). Whereas other clusters were surveyed in a region of radius 1.5 $r_{A}$, the photographic plate size restricted survey of Abell 1060 to a region of radius somewhat less than one Abell radius (see \\S \\ref{platty}). \\citeN{r89}, hereafter R89, presents a catalogue of 581 galaxies in the cluster area, which contains a sample which is complete to the magnitude limit $V_{25}=16.65$, within 2\\degr \\/ of the cluster centre. This is a convenient complete sample for the present H$\\alpha$ survey, extending to a fainter apparent magnitude than the Zwicky Catalogue used to define samples for other clusters. Cluster properties are summarised in Table \\ref{Hydratab}. The (B1950.0) position of the central cluster galaxy NGC 3311 is given as the cluster centre in columns 2 and 3. The mean heliocentric radial velocity $\\langle v_{\\odot}\\rangle$ and velocity dispersion $\\sigma$ determined using $N_{gal}$ individual galaxy redshifts are given in columns 4, 5, and 6 \\cite{bird94}. The Abell richness class \\cite{aco} is given in column 7, the Bautz-Morgan and Rood-Shastry type classes are given in columns 8 and 9 respectively \\cite{bm,strood}. \\begin{table*} \\centering \\caption{\\label{Hydratab} Cluster properties} \\begin{tabular}{ll@{\\hspace{1.0ex}}l@{\\hspace{1.0ex}}ll@{\\hspace{1.0ex}}l@{\\hspace{1.0ex}}lcccccc} \\\\ \\hline \\\\ Name & \\multicolumn{6}{c}{R.A.~~(1950.0)~~Dec.} & $\\langle v_{\\odot}\\rangle$ & \\(\\sigma\\) & $N_{gal}$ &Richness&\\multicolumn{2}{c}{Type Class} \\\\ \\cline{12-13} &\\multicolumn{3}{c}{l}&\\multicolumn{3}{c}{b}&km\\,s$^{-1}$&km\\,s$^{-1}$&&&B-M&R-S\\\\ \\\\ (1)&\\multicolumn{3}{c}{(2)}&\\multicolumn{3}{c}{(3)}&(4)&(5)&(6)&(7)&(8)&(9)\\\\ \\\\ \\hline \\\\ Abell 1060 &10$^{{\\rm h}\\hspace{-0.85ex}}$ &34$^{{\\rm m}\\hspace{-1.05ex}}$ &21\\fs 6&-27$^{\\circ\\hspace{-0.85ex}}$ &16$^{\\prime\\hspace{-0.65ex}}$ &05$^{\\prime\\prime\\hspace{-0.85ex}}$&3697 & 630 & 132 & 1 & III & C\\\\ (Hydra I)&\\multicolumn{3}{c}{269\\fdg6}&\\multicolumn{3}{c}{26\\fdg49}\\\\ \\\\ \\hline \\\\ \\end{tabular} \\end{table*} The observations and data reduction are described in \\S \\ref{obses}. Details of the observational technique are given in \\S \\ref{platty}, and of the process of identifying the emission-line galaxies in \\S \\ref{ident}, where a table of the detected emission-line galaxies (ELGs) is given. Measurements of radial velocities for the detected emission-line galaxies are presented in \\S \\ref{twoplate}, and those of \\mbox{H$\\alpha$+N[II]} equivalent widths and fluxes in \\S \\ref{ews}, where measured \\mbox{H$\\alpha$+N[II]} fluxes are also converted into effective star formation rates. A comparison of detected cluster emission-line galaxies in Abell 1060 with field galaxies and detected emission-line galaxies in other clusters is given in \\S \\ref{frac}. Notes on individual galaxies are given in \\S \\ref{indiv}. Finally, we present a brief discussion of our results in \\S \\ref{discus}. ", "conclusions": "" }, "9804/astro-ph9804030_arXiv.txt": { "abstract": "We present the first radio observations of a sample of 13 optically and IR bright southern hemisphere classical Be stars made from the Australian Telescope Compact Array at 3.5cm and 6.3cm simultaneously. One star, $\\delta$ Cen was detected at 3.5cm and a second, $\\mu$ Cen was also thought to have been detected; further observations of this source are required to confirm this detection. No sources were detected at 6.3cm, although $\\delta$ Cen was previously detected at this wavelength by other observers at a higher flux than our detection limit. The radio observations show that the spectral energy distribution undergoes a turnover between the far IR and radio wavelengths, as was seen in previous studies. Likewise we find no simple correlation between far IR and radio flux. Lower limits to the outer disc radius were found to be of the order of few hundred solar radii; of the order of those found previously by Taylor et al. ", "introduction": "Classical Be stars are defined as non supergiant B stars that have, or have had Balmer lines in emission. They are further characterised by the presence of a continuum excess, arising from free free and bound free emission from a stellar wind. Comparison of optical and UV spectra show that two different wind regimes must coexist. Consequently a high velocity component responsible for the high excitation lines visible in the UV, and a denser component that produces the near IR continuum excess and the optical emission lines was proposed to explain the observations. That the denser component was concentrated in the equatorial plane has long been suspected; recent interferometric data shows that the the envelopes around Be stars are non spherical (Dougherty \\& Taylor 1992; Quirrenbach et al 1994; Stee et al 1995). However, only a few of the brightest systems are ammeanable to an interferometric approach. Consequently, other approaches to the study of Be star circumstellar discs have been attempted. One such approach was long wavelength (mm-radio) flux measurement. When combined with near IR photometry modeling of the spectral energy distribution (SED) leads to a profile of the base density and density gradient (and hence a radial velocity law) within the circumstellar disc. However, observations of Be stars at mm and radio wavelengths showed that they were much fainter than implied by a simple extrapolation of their near IR and IRAS fluxes (e.g. Taylor et al 1990; henceforth Ta90), indicating a change in the ion density gradient within the disc. Several explanations were advanced to explain this, the most favourable being a change in disc opening angle or re-acceleration of material at large radii. Other possibilities exist, such as recombination at large radii or a truncation of the disc; see Waters et al. (1991) and Ta90 for a thorough discussion of all the scenarios. Given the paucity of detections in the northern hemisphere (only 6 Be stars have been detected), and the lack of multiple observations it is difficult to quantify the behaviour of the continuum excess at long wavelengths. Because of this shortfall we made observations of a sample of 13 bright southern hemisphere Be stars from the Australian Telescope Compact Array (ATCA) in 1997 April/May in an attempt to increase the sample size. ", "conclusions": "As a result of observations of a sample of 13 IR bright southern Be stars we have identified $\\delta$ Cen as a radio source, which may also be variable. A second star, $\\mu$ Cen, is also though to be a radio emitter, although further observations are needed to confirm this. This brings the total number of Be stars detected at radio wavelengths to eight (inclusive of $\\mu$ Cen). We confirm earlier results of Ta90 that demonstrate a turnover in the spectral energy distribution between the far IR and radio wavelengths. We find no compelling evidence for a direct correlation between stellar or far IR luminosity and radio flux. Clearly further observations of Be stars at higher sensitivities are required before such a correlation can be confirmed or rejected. Measurement of the minimum outer disc radius reveals that the circumstellar discs extend to several hundred solar radii, again reproducing the results of Ta90." }, "9804/astro-ph9804340_arXiv.txt": { "abstract": "In the process of searching the Hubble Space Telescope archive, we have serendipitously discovered three populous Large Magellanic Cloud (LMC) clusters with ages that place them in the LMC `age gap.'. These clusters - NGC 2155, SL663, and NGC 2121 - turn out to have $[Fe/H]$$\\sim$ --1.0 and ages of $\\sim$4 Gyr. This puts them in the age gap between the intermediate-age LMC clusters, the oldest of which are $\\sim$2.5 Gyr old, and ESO121-SC03, which has an age of $\\sim$ 9 Gyr. The addition of these three clusters to the LMC age - metallicity relation has reduced the discrepancy between the age distribution of the LMC clusters and the field stars. Furthermore, it indicates that searches to find more clusters older than $\\sim$2.5 Gyr in the LMC are crucial to a better understanding of its global star formation history. ", "introduction": "The chemical enrichment/star formation history (SFH) is an identifying feature of every self-gravitating stellar system. One manifestation of this is the relation between age and metallicity among the star clusters in a given galaxy. Empirical information on how cluster age and metal abundance correlate provides an important clue that will eventually allow us to understand how star formation (and hence chemical enrichment) proceeds under a variety of potentially influential circumstances. It is for this reason that we strive to better define the age - metallicity relations of the cluster and field populations in galaxies. Nearby galaxies are no exception, especially the Large and Small Magellanic Clouds (L/SMC) which have provided numerous puzzles and challenges for theorists. One of the most persistent of these has been the `age gap' seen between the intermediate-age clusters ($t \\lea 2.5$ Gyr; see Sec. 4.1 for a justification of this limit) and the old clusters ($t \\gea 13$ Gyr) in the LMC (Geisler et al. \\cite{geisea1997}). The LMC age gap is also a metallicity gap in the sense that the intermediate-age clusters have $\\langle$$[Fe/H]$$\\rangle$$\\sim$$-0.5$, while the old clusters are closer to $\\langle$$[Fe/H]$$\\rangle$$\\sim$$-2$. There is only one cluster that is known to lie in the gap - ESO121-SC03 with an age of $\\sim$9 Gyr and $[Fe/H]\\sim-1$. In contrast, recent studies based on Hubble Space Telescope (HST) data of the LMC field stars tell a signficantly different story (see Geha et al. \\cite{gehea1998} for a review). While there are variations that depend on position in the LMC (Vallenari et al. \\cite{valea1996}), the global star formation rate has been fairly constant for {\\it most} of the LMC's history. However, sometime between 2 and 4 Gyr ago, a burst of star formation occurred producing the present population of young to intermediate age field stars and clusters (Bertelli et al. \\cite{bertea1992}; Vallenari et al. \\cite{valea1996}; Gallagher et al. \\cite{gallea1996}; Holtzman et al. \\cite{holtzea1997}). Furthermore, the models produced by Geha et al. (\\cite{gehea1998}) suggest that roughly half of the LMC field stars are older than 4 Gyr. What this implies is that there should be many star clusters with ages between $\\sim$2.5 Gyr and $\\sim$13 Gyr, which were formed contemporaneously with the field stars. The obvious question of course is: where are these `age gap' clusters? A number of investigators have undertaken surveys to find clusters in the LMC age gap (Da Costa \\cite{dac1991}; Geisler et al. \\cite{geisea1997}, and references therein). The overwhelming conclusion has been that the LMC contains only one cluster with an age between 2.5 Gyr and 13 Gyr (i.e. ESO121-SC03). There is the possibility that some clusters did exist in the age gap sometime in the past, and that these have either dissolved into the LMC field or been stripped off (Olszewski \\cite{olsz1993}). However, it is difficult to see how this process can preferentially affect only clusters formed during a given epoch, unless the mass spectrum of density fluctuations producing the gap clusters was somehow different from that which produced the other clusters in the LMC. The possible reasons for this are not obvious. While understanding the SFH of the LMC is extremely important, this was not our original aim when we began this study. Initially, we were searching the Hubble Space Telescope (HST) archive looking for intermediate age clusters that display the red giant branch (RGB) Bump in their color-magnitude diagrams (CMD). We did find one such cluster (NGC 411; see Alves \\& Sarajedini \\cite{alsa1998}). However, more importantly for the present work, we discovered three clusters whose CMDs exhibit significant numbers of stars in the Hertzsprung gap, indicative of ages older than $\\sim$2.5 Gyr. These clusters are NGC 2155, NGC 2121, and SL663, and the next section describes the observations of these clusters and the data reduction. Section 3 presents the CMDs while Section 4 discusses the ages yielded by these CMDs. The implications of these results are detailed in Section 5. ", "conclusions": "The resulting relation between age and metallicity for the LMC star clusters is shown in Fig. 11. The open symbols are the ages and abundances from the work of Geisler et al. (\\cite{geisea1997}) supplemented by additional clusters from Bica et al. (\\cite{bicaea1998}) and the values for NGC 2193, Hodge 4, and ESO121-SC03 from this paper. The filled square represents the location of our three `age gap' clusters. Clearly, the clusters NGC 2155, NGC 2121, and SL 663 do indeed fall in the age gap between $\\sim$2.5 Gyr and $\\sim$13 Gyr. The reader should keep in mind, however, that if the metallicities of these clusters are higher than the values we have adopted herein, their ages will be correspondingly younger (see Sec. 4.1). As such, they may eventually be considered as belonging to the old-age tail of the IAC distribution. Future spectroscopic abundance measurements will shed more light on this. In any event, the addition of these three clusters to the age - metallicity relation of the LMC has not eliminated the discrepancy between the cluster age distribution and that of the field stars. If there are no more clusters to be discovered in the gap, then we will require some explanation for why the clusters and the field stars exhibit such differing SFHs. However, what is {\\it more} likely is that there are as yet unstudied clusters in the LMC that will further fill in the age gap. Future ground-based and HST photometric surveys may reveal more such clusters. For the present paper, we have utilized archival HST/WFPC2 images of LMC populous clusters to show that there are at least three clusters in the LMC age gap - NGC 2155, SL663, and NGC 2121. These clusters have $[Fe/H]$$\\sim$ --1.0 and ages of $\\sim$4 Gyr. The addition of these three clusters to the LMC age - metallicity relation is the first step in reducing the significant difference between the inferred SFHs of the LMC clusters and the field stars. This strongly indicates that searches to find more clusters older than $\\sim$2.5 Gyr in the LMC are crucial to a better understanding of its global SFH." }, "9804/astro-ph9804083_arXiv.txt": { "abstract": "Defect models have recently been declared dead\\cite{watson97}, because they predict microwave background and matter fluctuations grossly out of line with what we see. In this talk we apply the fact that many defects are automatically destroyed at the time of radiation-matter transition, thus resurrecting the defects model. Moreover, the resurrected version predicts a cosmological constant, explains the apparent excess of hot clusters and the non-Gaussianity observed in galaxy surveys. If this model is correct, then the MAP and PLANCK missions will not measure what people expect them to (oscillations); rather, they will measure a broad hump. ", "introduction": " ", "conclusions": "" }, "9804/astro-ph9804169_arXiv.txt": { "abstract": "Surface brightness fluctuations (SBFs) are much brighter in the infrared than they are at optical wavelengths, making it possible to measure greater distances using IR SBFs. We report new \\Kp\\ (2.1\\micron) SBF measurements of nine galaxies in the Fornax and Eridanus clusters using a 1024$^2$-pixel HgCdTe array. We used improved analysis techniques to remove contributions to the SBFs from globular clusters and background galaxies, and we assess the relative importance of other sources of residual variance. We applied the improved methodology to our Fornax and Eridanus images and to our previously published Virgo cluster data. Apparent fluctuation magnitudes were used in conjunction with Cepheid distances to M31 and the Virgo cluster to calibrate the \\Kp\\ SBF distance scale. We find the absolute fluctuation magnitude $\\MKp\\,{=}\\,{-}5.61\\,{\\pm}\\,0.12$, with an intrinsic scatter to the calibration of 0.06 mag. No statistically significant change in \\MKp\\ is detected as a function of \\v-i. Our calibration is consistent with simple (constant age and metallicity) stellar population models. The lack of a correlation with \\v-i\\ in the context of the stellar population models implies that elliptical galaxies bluer than $\\v-i\\,{=}\\,1.2$ have SBFs dominated by younger (5--8 Gyr) populations and metallicities comparable to redder ellipticals. Significant contributions to the SBFs from anomalous populations of asymptotic giant branch stars are apparently uncommon in giant ellipticals. \\Kp\\ SBFs prove to be a reliable distance indicator as long as the residual variance from globular clusters and background galaxies is properly removed. Also, it is important that a sufficiently high signal-to-noise ratio be achieved to allow reliable sky subtraction because residual spatial variance can bias the measurement of the SBF power spectrum. ", "introduction": "Because the light from a distant galaxy comes from discrete but unresolved stars, Poisson statistics lead to mottling of the galaxy's otherwise smooth surface brightness profile. Surface brightness is independent of distance, but the amplitude of the surface brightness fluctuations (SBFs) is not. As the distance ($d$) to a galaxy increases, the number of stars ($n$) in a given resolution element increases as $d^2$, but the observed flux ($f$) from each star is reduced by $d^{-2}$, making surface brightness ($nf$) independent of distance. On the other hand, the rms variation in observed flux from region to region is $n^{1/2}f$, which scales as $d^{-1}$. The variance in surface brightness $nf^2$ normalized by the mean galaxy brightness $nf$ decreases with distance; distant galaxies appear smoother than nearby galaxies. Because the variance in surface brightness is proportional to the second moment of the stellar luminosity function, it is dominated by luminous red giant stars. Although individual stars are not resolved, measuring SBFs probes the stellar population of the galaxy directly. With good theoretical models for stellar populations, the absolute magnitude of SBFs can be calculated, allowing a direct determination of the distance that is independent of the global dynamics or the environment of the galaxy (to the extent that the stellar populations in old stellar systems are independent of these variables). Alternatively, measuring SBFs in galaxies with known distances provides insight into stellar populations, allowing comparison with stellar evolution models and providing an empirical calibration of the SBF distance scale. Good descriptions of the theory and practice of using SBFs as a distance measurement tool and stellar population probe can be found in several papers by Tonry and collaborators (Tonry et al. 1997; Jensen, Luppino, \\& Tonry 1996, hereafter JLT; Tonry 1991; Tonry, Ajhar, \\& Luppino 1990; Tonry \\& Schneider 1988), in papers by Sodemann \\& Thomsen (1995, 1996), and in the review by Jacoby et al. (1992). The first $K$-band SBF studies are described by JLT, Luppino \\& Tonry (1993), and Pahre \\& Mould (1994). J. Tonry and coworkers have completed an extensive survey of $I$-band SBF distances in a sample of several hundred early-type (dynamically hot elliptical and S0) galaxies which is more than 50\\% complete out to $\\sim$2800 \\kms\\ (Tonry et al. 1997). They find that the $I$-band absolute fluctuation magnitude \\MI\\ is a linear function of \\v-i\\ and has a universal zero point. Their $I$-band calibration, which is based empirically on Cepheid distances, is in good agreement with Worthey's (1993a,b; 1994) simple stellar populations models. Worthey's models predict that the effects of age and metallicity in the $I$ band are largely degenerate, so that \\MI\\ may be calibrated using a single parameter such as the \\v-i\\ color. The resulting intrinsic scatter of the $I$-band SBF distance scale is of order 0.07 mag. The purpose of the current study is to examine the behavior of the SBF calibration in the near-IR \\Kp\\ band, where Worthey's models predict a much weaker dependence on \\v-i, but potentially larger scatter as the effects of age and metallicity are no longer degenerate. Stellar surface brightness fluctuations are very red, since they are dominated by luminous red giant stars. The advantages of observing IR SBFs are clear: fluctuations are ${\\sim}33$ times brighter at $K$ than at $I$, making them observable to greater distances. The SBF amplitude is inversely proportional to the seeing FWHM, and the seeing is typically much better in the IR than at optical wavelengths. Fluctuations are also red compared to the globular cluster (GC) population, so the contrast between stellar SBFs and GCs is higher at $K$. Finally, stellar population models predict that $K$-band SBF magnitudes have a much weaker dependence on \\v-i\\ than at $I$, reducing or eliminating the need to accurately measure the color of the galaxy (Worthey 1993a). Several studies have demonstrated the feasibility of measuring IR SBFs, including Luppino \\& Tonry (1993), Pahre \\& Mould (1994), and JLT. These papers report results for a rather limited set of galaxies in the Local Group (LG) and the Virgo Cluster, but the consistency in the measured calibration of \\MK\\ is encouraging ($\\MKp\\,{=}\\,{-}5.61\\,{\\pm}\\,0.16$ Luppino \\& Tonry, $M_{K_{sh}}\\,{=}\\,{-}5.77\\,{\\pm}\\,0.18$ Pahre \\& Mould, $\\MKp\\,{=}\\,{-}5.62\\,{\\pm}\\,0.29$ JLT). All three of these papers report absolute fluctuation magnitudes that are consistent with Worthey's (1994) predictions based on simple stellar population models. While the mean \\MKp\\ of JLT and Pahre \\& Mould (1994) are quite consistent, differences in apparent fluctuation magnitudes for individual galaxies are larger than the stated uncertainties allow. JLT discuss several possible reasons for the disagreement in fluctuation magnitudes and showed that residual variance resulting from spatial variations in the detector sensitivity and dark current can contribute significantly to the power spectrum in low signal-to-noise ratio ($S/N$) observations. A residual pattern of only 0.1\\% of the sky level can significantly change the fluctuation magnitude measured. To use IR SBFs as a reliable distance indicator, it is critical that the observations be of sufficiently high $S/N$ ratio to avoid biases from residual variances. JLT also demonstrated that properly subtracting the sky and carefully sampling the point spread function (PSF) are crucial to accurately measuring the fluctuation amplitude. An uncertainty in the background subtraction of 2\\% of the sky level can dominate the uncertainty in the fluctuation magnitude. Our Virgo sample showed a large dispersion (0.29 mag) due to the depth of the cluster and observational errors. Pahre \\& Mould (1994) also addressed the dispersion in fluctuation magnitude in a similarly-sized sample of Virgo ellipticals. Clearly a larger sample of galaxies is needed to calibrate the IR SBF distance scale. To better understand the effects stellar populations have on the fluctuation amplitude, we observed five galaxies in the Fornax cluster. Fornax is ideal for several reasons: first, it is much more centrally concentrated than the Virgo cluster. The reduced dispersion in distances allows us to better quantify and understand the uncertainty in our SBF measurements. Second, Fornax contains a large number of giant elliptical galaxies with a wide range of \\v-i\\ colors, metallicities, and globular cluster populations. The Virgo galaxies observed by JLT and Pahre \\& Mould (1994) spanned a very limited range in \\Mg2\\ index to avoid stellar population variations which may affect the calibration of \\MK. We now extend the sample to calibrate \\Kp\\ fluctuation magnitudes across a wider range of metallicities and \\v-i\\ colors to compare with theoretical stellar population models. Finally, Cepheid distances to several Virgo cluster galaxies and to the Fornax spiral NGC~1365 have been measured using the Hubble Space Telescope (HST). We can empirically anchor our \\Kp-band SBF calibration both to the Cepheid and the $I$-band SBF distance scales. This paper reports results from our observations of four Eridanus cluster elliptical galaxies in addition to the Fornax galaxies. Eridanus is not as compact a cluster as Fornax, but we can still compare our IR SBF results with the $I$-band results from Tonry (1997). Finally, we return to our SBF measurements for the seven Virgo ellipticals discussed by JLT and apply the improved techniques we describe in this paper. We present the updated results for these galaxies in Section~\\ref{virgorevisited}, along with new observations of NGC~4365. ", "conclusions": " $K$-band SBFs can be measured reliably and used to determine distances to early-type galaxies. Because SBFs are relatively bright in the IR and the seeing is typically very good, integration times can be quite modest. However, observations must be sufficiently deep to adequately sample the GC luminosity function and remove sources of residual variance. Low-$S/N$ ratio measurements are unreliable. Optical images can be used to identify and remove GCs and background galaxies that are not detected in the IR, improving the IR SBF measurement. We empirically calibrated the \\Kp\\ SBF distance scale using Cepheid distances to M31 and the Virgo cluster. The absolute \\Kp\\ fluctuation magnitude is $\\MKp\\,{=}\\,{-}5.61\\,{\\pm}\\,0.06$ (statistical error) with a total uncertainty of 0.12 mag. No significant change in \\MKp\\ with \\v-i\\ was observed over the range in color spanned by the galaxies in this sample. Accurate color measurements are not required to measure the \\Kp\\ SBF distance to a galaxy. \\Kp\\ SBF magnitudes are consistent with predictions from simple stellar population models. The lack of a correlation in \\MKp\\ with \\v-i\\ is best explained by a spread in ages among the galaxies observed. The redder ellipticals are consistent with 12 to 17 Gyr stellar population models, while the bluer galaxies in our sample must have younger 5 to 8 Gyr populations. Metallicities appear to vary less than ages, with the typical galaxy having [Fe/H]\\,=\\,${-}0.25$. The stellar population models show that the age-metallicity degeneracy is broken with $K$-band SBFs, allowing one to distinguish between old, metal-poor and young, metal-rich populations. Examining the radial variation in IR SBFs will help distinguish between different galaxy formation scenarios. Our observations do not agree with the relationship between \\MKp\\ and the \\Mg2\\ index predicted by the stellar population models." }, "9804/astro-ph9804263_arXiv.txt": { "abstract": "The high redshift radiogalaxy 1243+036 ($z=3.6$) presents an asymmetric \\lya\\ profile of FWHM 1550 \\kms\\ as measured by van~Ojik \\etal\\ We propose that the blue asymmetry in the \\lya\\ profile is not due to narrow absorption dips but consists of narrow emission peaks. We interpret the blueshifted peaks near $-1130, -850$ and $-550 \\kms$ (relative to the peak of full profile) as being the result of Fermi acceleration of \\lya\\ produced by jet-induced star formation in the wake of a 300~\\kms\\ shock. This shock would be caused by the deflection of the radio-jet at the observed position of the radio bend which also coincides spatially with the excess \\lya\\ emission reported by van~Ojik \\etal\\ ", "introduction": "\\label{intro} Lyman $\\alpha$ is the strongest line observed in very high redshift radio galaxies (HZRG). Although the brightness of \\lya\\ peaks near or at the nuclear position, most of the emission is spatially resolved with the fainter emission extending up to radii 40--130\\,kpc. The \\lya\\ profile in HZRG is characterized by a FWHM in the range 700--1600\\kms\\ (van~Ojik 1995 and references therein). The intermediate resolution study of \\lya\\ profiles carried out by van~Ojik \\etal\\ (1997) has revealed the presence of troughs which are well explained by \\hi\\ gas absorption present in the environment of the parent radio galaxy. In their sample, however, the HZRG 1243+036 ($z=3.6$) may call for a different interpretation, namely that the broad profile presents true narrow emission features in between the narrow `dips' which van~Ojik \\etal\\ (1996: vO96) have interpreted as absorption features. The repeated scattering of the resonance \\lya\\ line across a shock discontinuity was shown by Neufeld \\& McKee (1988: NM88) to result in a systematic blueshift of \\lya. In this paper, we develop in more detail the Fermi acceleration model and propose that the narrow features observed by vO96 correspond to a small number of across-shock scatterings. ", "conclusions": "" }, "9804/astro-ph9804325_arXiv.txt": { "abstract": "Recent discoveries by the {\\it Rossi X-Ray Timing Explorer} indicate that most of the rapidly accreting ($\\dot M \\gtrsim 10^{-11} M_\\odot \\ {\\rm yr}^{-1}$) weakly magnetic ($B\\ll 10^{11} \\ {\\rm G}$) neutron stars in the Galaxy are rotating at spin frequencies $\\nu_s \\gtrsim 250 \\ {\\rm Hz}$. Remarkably, they all rotate in a narrow range of frequencies (no more than a factor of two, with many within 20\\% of 300 Hz). I suggest that these stars rotate fast enough so that, on average, the angular momentum added by accretion is lost to gravitational radiation. The strong $\\nu_s$ dependence of the angular momentum loss rate from gravitational radiation then provides a natural reason for similar spin frequencies. Provided that the interior temperature has a large scale asymmetry misaligned from the spin axis, then the temperature sensitive electron captures in the deep crust can provide the quadrupole needed ($\\sim 10^{-7} M R^2$) to reach this limiting situation at $\\nu_s\\approx 300$ Hz. This quadrupole is only present during accretion and makes it difficult to form radio pulsars with $\\nu_s>(600-800) \\ {\\rm Hz}$ by accreting at $\\dot M \\gtrsim 10^{-10} M_\\odot \\ {\\rm yr^{-1}}$. The gravity wave strength is $h_c\\sim (0.5-1) \\times 10^{-26}$ from many of these neutron stars and $>2\\times 10^{-26}$ for Sco X-1. Prior knowledge of the position, spin frequency and orbital periods will allow for deep searches for these periodic signals with gravitational wave interferometers (LIGO, VIRGO and the ``dual-recycled'' GEO 600 detector) and experimenters need to take such sources into account. Sco X-1 will most likely be detected first. ", "introduction": "The launch of the {\\it Rossi X-Ray Timing Explorer} (RXTE) has allowed for the discovery of fast quasi-periodic variability from many rapidly accreting ($ \\dot M\\gtrsim 10^{-11} M_\\odot \\ {\\rm yr^{-1}}$) neutron stars. These observations strongly suggest that these neutron stars (NSs) are rapidly rotating, as predicted by those scenarios connecting the millisecond radio pulsars to this accreting population (see Bhattacharya 1995 for an overview). Strohmayer et al. (1996) were the first to detect nearly coherent $\\nu_B=363$ Hz oscillations during type I X-ray bursts from the low accretion rate ($\\dot M< 10^{-9} M_\\odot \\ {\\rm yr^{-1}}$) NS 4U~1728-34. Pulsations were detected in six of the eight bursts analyzed at that time. In addition, two high frequency quasi-periodic oscillations (QPOs) were seen in the persistent emission. These changed with accretion rate, but maintained a fixed difference frequency of $\\nu_d\\approx 363 $ Hz, identical to the period seen during the bursts. The detection of two drifting QPO's (in the persistent emission) separated by a fixed frequency identical to that seen in the bursts naturally leads to beat frequency models (Strohmayer et al. 1996; Miller, Lamb, \\& Psaltis 1998). The difference frequency is presumed to be the NS spin frequency, $\\nu_s$, whereas the upper frequency has different origins in different models (see van der Klis 1998 for a summary). In addition, the temporal behavior of the periodic oscillations both during the rise of the bursts (Strohmayer, Zhang, \\& Swank 1997b) and in the cooling tails (Strohmayer et al. 1997a) are most easily explained in terms of rotation. There are six NSs with measured periodicities during Type I X-ray bursts (see Table 1). Both the difference frequencies ($\\nu_d$) and the burst frequencies ($\\nu_B$) are in a narrow range, from 260 to 589 Hz. For two objects (KS 1731-260 and 4U 1636-53) the difference frequencies are one-half the burst values. Which value is $\\nu_s$ is not resolved. There are also many NSs that accrete at higher rates and are not regular Type I X-ray bursters. Many of these objects, notably the ``Z'' sources, also show drifting QPO's at fixed separation, again with a similarly narrow frequency range (roughly 250-350 Hz). Beat-Frequency like models are also applied to these observations so as to infer $\\nu_s$. The applicability of such a model is less clear when the difference frequency is not constant (Sco X-1, van der Klis et al. 1997; 4U 1608-52, Mendez et al. 1998). If accreting matter always arrives with the specific angular momentum of a particle orbiting at the NS radius ($R=10 R_6 {\\rm km}$), then it only takes $\\sim 10^7 \\ {\\rm yrs}$ of accretion at $\\dot M\\approx 10^{-9} M_\\odot \\ {\\rm yr^{-1}}$ for a $M=1.4M_{1.4} M_\\odot$ NS to reach $\\nu_s=50$ Hz from an initially low frequency. It is thus remarkable that these NSs are all rotating at nearly the same rate. White and Zhang (1997) argued that this similarity arises because these NSs are magnetic and have reached an equilibrium where the magnetospheric radius equals the co-rotation radius. This requires an intrinsic relation between their magnetic dipoles, $\\mu_b$, and $\\dot M$ so that they all reach the same rotational equilibrium (most likely $\\mu_b\\propto \\dot M^{1/2}$) and a way of hiding the persistent pulse typically seen from a magnetic accretor. My alternative explanation for these spin similarities is that gravitational wave (GW) emission has started to play an important dnrole. If the NS has a misaligned quadrupole moment, $Q$, then the strong spin frequency dependence of GW emission defines a critical frequency beyond which accretion can no longer spin-up the star. Such a NS will radiate energy via GW's at the rate $\\dot E=32 GQ^2\\omega^6/5 c^5$, where $\\omega=2\\pi \\nu_s$, and lose angular momentum at the rate $N_{gw}=\\dot E/\\omega$. Balancing this spin-down torque with the characteristic spin-up torque from time-averaged accretion, $N_a\\approx \\dot M(GMR)^{1/2}$, gives the $Q$ needed so as to make the critical frequency 300 Hz, \\begin{equation}\\label{eq:qneed} Q\\approx 4.5 \\times 10^{37} \\ {\\rm g \\ cm^2}\\left(\\dot M\\over 10^{-9} \\ {\\rm M_\\odot \\ yr^{-1}}\\right)^{1/2}\\left(300 \\ {\\rm Hz}\\over \\nu_s\\right)^{5/2}, \\end{equation} or $<10^{-7}$ of the NS moment of inertia, $I\\approx 10^{45} \\ {\\rm g \\ cm^2}$. The similarities in $\\nu_s$ may then arise because of the weak dependencies of the critical frequency on $Q$ and $\\dot M$. What is the source of the misaligned quadrupole? Wagoner (1984) argued that accreting NSs would get hung-up at spin frequencies where the Chandrasekhar-Friedman-Schutz (CFS) instability sets in. However, Lindblom (1995) and Lindblom \\& Mendell (1995) have shown that the star needs to be very near the breakup frequency ($\\nu_s \\gtrsim \\ {\\rm kHz}$) for such an instability to occur, even for the core temperatures $T_c=(1-3)\\times 10^8 \\ {\\rm K}$ of rapidly accreting NSs (Ayasli \\& Joss 1978; Brown \\& Bildsten 1998). The spin frequencies for these NSs are too slow for such an instability. I present in \\S 2 a new source for lateral density variations in an accreting NS; electron captures (hereafter EC) in the crust. The constant compression of the crust forces nuclei to undergo EC when the electron Fermi energy, $E_F$, is high enough to make a transition. However, the crust is hot enough in a rapidly accreting NSs to make the EC rates temperature sensitive. Hotter regions then capture at lower pressures, so that the density jump associated with the EC is at a higher altitude in the hotter parts of the crust. Moderate lateral temperature variations then lead to density variations large enough to generate the required $Q$. This outcome is independent of the particular source of the temperature variations. One possible cause for a $T$ asymmetry relative to the spin axis is a weak magnetic field. I conclude in \\S 3 by finding the GW signal strength and estimating detection. ", "conclusions": "" }, "9804/astro-ph9804113_arXiv.txt": { "abstract": "\\vspace{-3ex} We present a numerical simulation of the dynamical collapse of a nonrotating, magnetic molecular cloud core and follow the core's evolution through the formation of a central point mass and its subsequent growth to a $1~\\msol$ protostar. The epoch of point-mass formation (PMF) is investigated by a self-consistent extension of previously presented models of core formation and contraction in axisymmetric, self-gravitating, isothermal, magnetically supported interstellar molecular clouds. Prior to PMF, the core is dynamically contracting and is not well approximated by a quasistatic equilibrium model. Ambipolar diffusion, which plays a key role in the early evolution of the core, is unimportant during the dynamical pre-PMF collapse phase. However, the appearance of a central mass, through its effect on the gravitational field in the inner core regions, leads to a ``revitalization'' of ambipolar diffusion in the weakly ionized gas surrounding the central protostar. This process is so efficient that it leads to a decoupling of the field from the matter and results in an outward-propagating hydromagnetic C-type shock. The existence of an ambipolar diffusion-mediated shock of this type was predicted by Li \\& McKee (1996), and we find that the basic shock structure given by their analytic model is well reproduced by our more accurate numerical results. Our calculation also demonstrates that ambipolar diffusion, rather than Ohmic diffusivity operating in the innermost core region, is the main field decoupling mechanism responsible for driving the shock after PMF. The passage of the shock leads to a substantial redistribution, by ambipolar diffusion but possibly also by magnetic interchange, of the mass contained within the magnetic flux tubes in the inner core. In particular, ambipolar diffusion reduces the flux initially threading a collapsing $\\sim 1~\\msol$ core by a factor $\\simgt 10^3$ by the time this mass accumulates within the inner radius ($\\simeq 7.3~{\\rm AU}$) of our computational grid. This reduction, which occurs primarily during the post-PMF phase of the collapse, represents a significant step towards the resolution of the protostellar magnetic flux problem. Our calculations indicate that a $1~\\msol$ protostar forms in $\\sim 1.5 \\times 10^5~\\rm{yr}$ for typical cloud parameters. The mass accretion rate is time dependent, in part because of the C-shock that decelerates the infalling matter as it propagates outward: the accretion rate rises to $\\simeq 9.4~\\msol~\\rm{Myr}^{-1}$ early on and decreases to $\\simeq 5.6~\\msol~{\\rm{Myr}}^{-1}$ by the time a solar-mass protostar is formed. The infalling gas disk surrounding the protostar has a mass $\\sim 10^{-2}~\\msol$ at radii $r \\simgt 500~\\rm{AU}$. A distinguishing prediction of our model is that the rapid ambipolar diffusion after the formation of a protostar should give rise to large ($\\simgt 1~{\\rm{km}}~{\\rm s}^{-1}$), and potentially measurable, ion--neutral drift speeds on scales $r \\simlt 200~\\rm{AU}$. The main features of our simulation, including the C-shock formation after PMF, are captured by a similarity solution that incorporates the effects of ambipolar diffusion (Contopoulos, Ciolek, \\& K\\\"{o}nigl 1997). ", "introduction": "It is generally accepted that most of the star-formation activity in our galaxy takes place through the gravitational collapse of molecular cloud cores (e.g., Mouschovias 1987; Shu, Adams, \\& Lizano 1987). It is, furthermore, believed that interstellar magnetic fields play a central role in this process in that their stresses are the dominant agent that acts against gravity to prevent, or delay, cloud collapse (e.g., Mouschovias 1978; McKee et al. 1993). This is embodied in the concept of a critical mass $M_{\\rm crit}$, which in general takes account of both the magnetic and the thermal pressure contributions to the support of the cloud, but which, in the case that magnetic stresses dominate, reduces to $M_{\\rm crit} \\approx 0.13 \\phiB/G^{1/2}$, where $G$ is the gravitational constant and $\\phiB$ is the magnetic flux that threads the cloud. Clouds whose mass $M$ exceeds $M_{\\rm crit}$ are ``supercritical'': they collapse on the free-fall timescale. In contrast, ``subcritical'' clouds ($M < M_{\\rm crit}$) can avoid collapse on the much longer ambipolar diffusion timescale. In the latter case, the neutrals gradually contract by diffusing inward through the ions and field, leaving behind a magnetically supported envelope and eventually forming a supercritical core that undergoes dynamical collapse (e.g., Mouschovias 1996). Because of the complexity of the problem --- it involves solving the full dynamical equations of a magnetized, multicomponent (neutrals, ions, electrons, as well as charged and neutral grains) fluid that evolves over many decades in size and density in a nonspherically symmetric manner (because of the presence of ordered magnetic fields and likely also rotation) --- much of the progress in this area has been accomplished through the use of numerical simulations. One of the main efforts to simulate core formation and contraction due to ambipolar diffusion in magnetically supported molecular clouds has been carried out by Mouschovias and coworkers (e.g., Fiedler \\& Mouschovias 1992, 1993; Ciolek \\& Mouschovias 1993, 1994, 1995, hereafter CM93, CM94, CM95, respectively; Basu \\& Mouschovias 1994, 1995a, b). These studies followed the evolution of the core over six decades in density up to central densities $\\sim 3 \\times 10^9~\\cc$, where the assumption of isothermality starts to break down because of radiative trapping (e.g., Gaustad 1963; Hayashi 1966). This assumption had been adopted in the interest of simplicity: sophisticated and computationally intensive numerical techniques are generally needed to calculate the thermal structure of the gas during the opaque phase of protostellar evolution (e.g., Larson 1969, 1972; Winkler \\& Newman 1980; Stahler, Shu, \\& Taam 1981; Boss 1984; Myhill \\& Kaula 1992; Myhill \\& Boss 1993). As a result of this restriction, the aforementioned calculations did not follow the collapse of the core to the time where a point mass -- a protostar -- is formed at the center, although they did obtain valuable information on the conditions leading to this critical event. In particular, by the time these simulations were terminated, the inner region of the core was collapsing dynamically and was characterized by neutral infall speeds $\\sim C$ (the isothermal speed of sound) and inward accelerations $\\simgt 0.3 |\\gr|$ [where $\\gr(r)$ is the local gravitational acceleration]. Furthermore, the thermal pressure, while remaining relatively unimportant in the envelope, came to exceed the magnetic pressure near the center. Basu (1997) derived a time-dependent, semianalytic solution that extended these ambipolar diffusion models up to the instant of point-mass formation (hereafter referred to as PMF \\footnote{Creation of a central point mass was commonly referred to in previous papers as ``core formation.'' In this paper we use the term ``point-mass formation'' so as not to confuse this process with the formation of an extended, magnetically supercritical, molecular cloud core.}). He found that ambipolar diffusion continues to gradually erode the retarding magnetic forces in the inner core, making the collapse increasingly more dynamical (and the thermal-to-magnetic pressure ratio in the inner core progressively larger) as PMF is approached. The diminution of magnetic forces in the innermost regions of a collapsing core just prior to PMF suggests that one could gain some insight into the protostar formation process from previous studies of PMF in {\\it nonmagnetic}, spherically symmetric, isothermal clouds. Analytic similarity solutions have uncovered two limiting behaviors: Penston (1969) and Larson (1969) found a solution in which, just before PMF, the infall speed approaches $\\sim 3.3~C$ at all radii while the density scales with radius as $r^{-2}$, resulting in a spatially uniform mass inflow rate $\\sim 29~C^3/G$ (where $G$ is the gravitational constant). Hunter (1977) extended this solution past PMF and showed that, immediately after the central mass is formed, the accretion rate onto the protostar increases to $\\sim 47~C^3/G$. In the other limit, Shu (1977) obtained a solution that is static prior to PMF (with the density distribution of a singular isothermal sphere, which also scales as $r^{-2}$) and that takes on an expansion-wave character (with a constant mass accretion rate $\\sim 0.98~C^3/G$ onto the protostar) following PMF. Hunter (1977) and Whitworth \\& Summers (1985) demonstrated that there are, in fact, infinitely many similarity solutions that span the range between the Larson-Penston and Shu results, with the nature of any given solution being determined by the initial configuration of the cloud and the conditions at its boundary. Numerical simulations carried out by Hunter (1977) and by Foster \\& Chevalier (1993) confirmed the dependence on the initial and boundary conditions. In particular, it was found that the behavior of the central regions of clouds that are initially marginally stable to collapse approximates that of the Larson-Penston solution at the PMF epoch, although it was determined that the mass accretion rate onto the protostar declines at later times. It was, however, also found that the post-PMF evolution of clouds that initially have more extended envelopes approximates that of the Shu solution at late times. Since the initial and boundary conditions of real clouds are expected to depend on the detailed configuration and evolution of the embedded magnetic field, it is clear that one needs to incorporate magnetic field effects into the collapse calculations to adequately model the formation of protostars. There have been several recent attempts to calculate PMF following the collapse of magnetic interstellar clouds. Although they have all contributed to our understanding of the processes involved, their results were hampered by the adopted assumptions or approximations. For example, Tomisaka (1996) modeled clouds that had equal thermal and magnetic energy densities, so that they were not primarily supported by magnetic fields. This means that his model clouds were magnetically supercritical. This assumption is at variance with H {\\sc I} and OH Zeeman measurements of magnetic field strengths in molecular clouds, which are consistent with models of magnetically subcritical clouds (Crutcher et al. 1993, 1994, 1996). Li \\& Shu (1997) modeled PMF in self-similar, magnetic cores. They assumed that cores immediately before PMF can be represented by hydrostatic configurations of singular isothermal disks and that the magnetic flux is frozen into the neutrals. These assumptions are inconsistent with the above-cited results of numerical simulations and semianalytic solutions of the collapse of magnetically supported molecular clouds that undergo ambipolar diffusion (as well as with the simulations of thermally supported spherical clouds), which have found that the inner core regions collapse dynamically as PMF is approached (see also \\S 3.2). As we show below, ambipolar diffusion, which plays a key role in bringing about the dynamical collapse, is generally important also {\\it after} PMF. Safier, McKee, \\& Stahler (1997) did include ambipolar diffusion in the modeling of magnetic cloud collapse and post-PMF evolution. However, their model is strictly spherical, and they neglected the effect of thermal pressure gradients in comparison with magnetic stresses. Furthermore, their formulation is quasistatic and does not incorporate the magnetic induction equation for the time evolution of the magnetic field. Li (1998) extended the Safier et al. model by including thermal-pressure and time-dependent terms and by adding the induction equation. This enabled him to follow the time evolution of his model cores (for $r > 150~\\rm{AU}$) even during the dynamical phases of the collapse. However, by retaining the spherical-symmetry assumption of Safier et al., his calculations were also unable to yield the geometry of the magnetic field lines. \\footnote{The same is true for the spherically symmetric self-similar model of a collapsing magnetic cloud devised by Chiueh \\& Chou (1994), which, however, does not include ambipolar diffusion. It should be noted that all models that assume spherical symmetry do not satisfy the solenoidal condition $\\nabla \\cdot \\Bvec =0$ on the magnetic field.} The importance of ambipolar diffusion after PMF can be inferred by comparing the ambipolar diffusion timescale $\\tad = r/\\vd$ (where $\\vd$ is the ion--neutral drift speed) and the gravitational contraction ($\\simeq$ free-fall) timescale $\\tgr = (r/|\\gr|)^{1/2}$ before and after PMF. In axisymmetric geometry, $\\tad/\\tgr \\simeq (\\tgr/\\tni) \\, {\\mu_ B}^2$ in the inner flux tubes of a supercritical core (e.g., Mouschovias 1991), where ${\\mu_B}(r)$ is the total mass-to-flux ratio at radius $r$ (in units of the critical value for gravitational collapse) and \\begin{equation} \\label{tnieq} \\tni = 1.4 \\left[1 + 0.067 \\frac{\\left(\\mHII /2~\\rm{a.m.u.}\\right)}{\\left(\\mi/30~\\rm{a.m.u.}\\right)}\\right] \\frac{1}{\\nni \\sigw} \\end{equation} is the neutral--ion collision time. In equation (\\ref{tnieq}), $\\nni$ is the ion density and $\\sigw$ is the average collisional rate between ions of mass $\\mi$ and neutral $\\HII$ molecules of mass $\\mHII$ ($\\simeq 1.7 \\times 10^{-9}~{\\rm{cm^3}}~{\\rm s}^{-1}$ for collisions between neutrals and $\\Mgp$ or $\\HCOp$ ions; McDaniel \\& Mason 1973); the factor 1.4 accounts for a 20\\% helium abundance by number. CM94 and CM95 found that, to first order for the late pre-PMF evolution of cores in disk-like clouds, the magnetic field $B \\approx 3\\, (r_0/r)~\\rm{mG}$ (where $r_0 = 40~\\rm{AU}$), $\\nni \\approx 0.1~\\cc$ (valid for neutral densities $\\nn \\simgt 10^7~\\cc$; see Figs. $2b$ and $4b$ in CM94), the vertical column density $\\sign \\approx 5\\, (r_0/r)~{\\rm g}~\\rm{cm}^{-2}$, and the total mass $M(r) \\approx 6 \\times 10^{-3} (r/r_0) \\msol$. For a disk-like cloud, the critical mass-to-flux ratio $(M/\\phiB)_{\\rm{d,crit}}= (4 \\pi^2 G)^{-1/2}$, where $G$ is the gravitational constant (Nakano \\& Nakamura 1978). Therefore, one finds $\\mu_B \\approx 2.7$, $|\\gr| \\approx G M(r)/r^2 \\approx 2 \\times 10^{-6} (r_0/r) ~{\\rm{cm}}~\\rm{s}^{-2}$, and $\\tni \\approx 7 \\times 10^9~\\rm{s}$, which yields $\\tgr \\approx 2 \\times 10^{10} (r/r_0) ~\\rm{s}$ and $\\tad/\\tgr \\approx 20\\, (r/r_0)$. It follows that $\\tad/\\tgr \\gg 1$ for $r \\gg 2~\\rm{AU}$. Hence, as has already been demonstrated in the past, ambipolar diffusion is ineffective as a dynamically collapsing core approaches PMF, and for $r \\gg 2~\\rm{AU}$ the magnetic flux can be considered frozen into the neutrals. (For reasons discussed in \\S 2.2, we do not consider $r \\simlt 5~\\rm{AU}$.) Turning now to the post-PMF epoch, when the central point mass comes to dominate the gravitational field in the innermost flux tubes, we use the relation $\\tad/\\tgr \\approx \\left(1- |\\an|/|\\gr|\\right)^{-1} (\\tgr/\\tni)$, where $\\an$ is the inward acceleration of the neutrals and $\\gr$ is the total gravitational acceleration (see eqs. [\\ref{rforceeq}] and [\\ref{tadeq}] in \\S 3.3). In this case, normalizing $r$ as before, $|\\gr| \\approx 4 \\times 10^{-4} (\\mcent/\\msol) (r_0/r)^2 {\\rm{cm}}~\\rm{s}^{-2}$. Substituting again $\\tni \\approx 7 \\times 10^9~{\\rm s}$ (corresponding to $\\nni \\approx 0.1~\\cc$) , the above ratio becomes $\\tad/\\tgr \\approx 0.2\\, (r/r_0)^{3/2}(\\msol/\\mcent)^{1/2}\\left(1-|\\an|/|\\gr|\\right)^{-1}$. For $|\\an|/|\\gr|$ in the range 0.2 -- 0.9, which corresponds to the period after PMF when the collapse becomes progressively more dynamical, one infers $\\tad/\\tgr \\approx (0.2 - 2) (r/r_0)^{3/2}(\\msol/\\mcent)^{1/2}$. Hence $\\tad \\simlt \\tgr$ for $\\mcent \\simgt 0.1 \\msol$ and $r \\simlt 50~\\rm{AU}$. [A similar result is obtained if one continues to use the pre-PMF relations and simply substitutes $\\mcent$ for $M(r)$.] This estimate indicates that {\\em decoupling of the gas and magnetic field by ambipolar diffusion should occur in the inner core regions after PMF}. The physical reason for this is that the strength of the gravitational field in the weakly ionized gas near the origin is greatly enhanced by the appearance and growth of a central point mass, causing the neutrals to fall in more rapidly while the plasma and magnetic field are left behind. (The same basic reason --- the appearance of a progressively growing free-fall zone around the origin following PMF --- is also the cause of the increase in the mass inflow rate into the center at that epoch first discovered in the above-referenced nonmagnetic collapse calculations.) The foregoing conclusion is verified by a detailed calculation in \\S~3.3, where we show that ambipolar diffusion after PMF is, in fact, so efficient that it effectively decouples the neutrals and magnetic field in the innermost core region, with dramatic consequences for the subsequent dynamical evolution of the core. An alternative, yet equivalent, analysis of the effectiveness of ambipolar diffusion following PMF, based on the scaling of magnetic forces (particularly the magnetic tension force) after PMF, is given in Appendix C. In this paper we present a detailed numerical simulation of point-mass formation in a nonrotating, magnetic, dynamically collapsing protostellar core, properly accounting for the effect of ambipolar diffusion. Unlike earlier studies, we use an initial state that is consistent with the realistic models of core formation and collapse, as presented earlier by Mouschovias and coworkers. As we noted above, the simulations carried out by that group were terminated at densities where the isothermality assumption started to become invalid because of radiative trapping. Although a proper treatment of radiative trapping is indispensable for a complete treatment of the star-formation process, one can adopt a simpler approach that circumvents this difficulty by removing the region of radiative trapping from the active computation mesh. This is justified by the fact that the region of radiative trapping (typically a few AU) is several orders of magnitude smaller than the characteristic core size ($r_{\\rm{core}} \\approx 0.1~\\rm{pc}$). This region can therefore be considered to be effectively point-like, and one can proceed to calculate the formation of a central point mass within a gravitationally collapsing core and its effect on the subsequent evolution of the core by retaining the isothermality assumption. In adopting this approach, we note that the assumption of isothermality was also employed in previous studies of protostar formation in nonmagnetic cloud cores as well as in the more recent attempts to model PMF in magnetic clouds. The plan of the paper is as follows. In \\S~2 we review the main characteristics of the pre-PMF core-evolution calculations of CM93, CM94, and CM95 and outline the model modifications that we have implemented to extend the simulations beyond PMF. In \\S~3 we present the results of our calculations. We consider a typical model and follow the evolution of all physical quantities of interest through the PMF epoch. We also describe the formation (due to ambipolar diffusion in the innermost core after PMF) of a C-type shock and consider its propagation through the collapsing core. The appearance of such a shock as a result of field--matter decoupling was first pointed out by Li \\& McKee (1996), who proposed that the relevant field decoupling mechanism was Ohmic dissipation in the innermost regions of the core. Our study, however, reveals that ambipolar diffusion occurring outside the region of Ohmic dissipation is the main cause of magnetic flux decoupling after PMF. In \\S~4 we discuss the structure of the ambipolar diffusion-mediated shock and show that our numerical calculations qualitatively reproduce the predictions of the simplified analytic model constructed by Li \\& McKee (1996). We present a quantitative comparison with their results and also address the issue of interchange instability in the post-shock region, first raised in their work, in light of our detailed computations. In that section we also discuss the observational implications of our simulation and briefly comment on the magnetic flux problem during star formation. Our results are summarized in \\S~5. ", "conclusions": "\\subsection{Observational Comparisons and Predictions} We may compare our result for the protostellar accretion rate during the PMF epoch with observations of star-forming cores. As shown in \\S~3.3, the accretion rate rises rapidly early on to $\\mcentdot \\simeq 9.4~\\msol~\\rm{Myr}^{-1}$ for $\\delt \\simlt 10^3~\\rm{yr}$ (see Fig. $2b$) and stays at this value up to the formation of the hydromagnetic shock. For $\\delt \\simgt 4 \\times 10^3~\\rm{yr}$ the shock is able to decelerate the infalling matter, and the accretion rate decreases to $\\mcentdot \\simeq 5.6~\\msol~\\rm{Myr}^{-1}$ by $\\delt \\simeq 1.5 \\times 10^5~\\rm{yr}$ (the time when $\\mcent = 1~\\msol$). Therefore $\\mcentdot$ decreases with increasing central mass (see Fig. $2c$). This is consistent with estimates of ages and accretion rates ($\\propto t_{\\rm{age}}^{-1}$) for young stellar objects, as deduced from evolutionary diagrams inferred from observations of Class 0 and Class I objects (e.g., Saraceno et al. 1996). In particular, the lifetimes of Class 0 objects were estimated in this way to be an order of magnitude shorter than those of Class I objects, providing evidence for a decrease in the protostellar accretion rate as an object evolves from a Class 0 source to a Class I source (e.g., Andr\\'{e} 1995; Ward-Thompson 1996). Another argument for a time-dependent accretion rate was given by Bontemps et al. (1996), who analyzed the observed CO momentum flux of several young stellar objects and found a noticeable decline in the CO flux with decreasing circumstellar envelope mass. They suggested that this is indicative of a decrease in the protostellar accretion rate (which they assumed to be proportional to the mass outflow rate) as an object evolves from Class 0 to Class I. It is also of interest to note the ${}^{13}{\\rm{CO}}(J=1-0)$ observations of infalling disks for the protostellar candidates HL Tauri (Hayashi, Ohashi, \\& Miyama 1993) and L1551-IRS5 (Ohashi et al. 1996). HL Tauri has a mass $\\sim 0.6~\\msol$ and a surrounding disk with radius $\\sim 1400~\\rm{AU}$ and mass $\\sim 0.03~\\msol$. From the observed kinematics Hayashi et al. derive an accretion rate $\\sim 9~\\msol~\\rm{Myr}^{-1}$ at $r \\sim 700~\\rm{AU}$. The embedded protostar L1551-IRS5 has a mass $\\sim 0.5~\\msol$ and is surrounded by a disk with radius $\\sim 700~\\rm{AU}$ and mass in the range $3.9 \\times 10^{-2} - 8.1 \\times 10^{-2}~\\msol$. Ohashi et al. deduce an accretion rate in the range $13 - 26~\\msol~{\\rm{Myr}}^{-1}$ at $r \\sim 600~\\rm{AU}$. These values are comparable to our model results for $\\delt_5 \\simlt \\delt \\simlt \\delt_6$ (corresponding to $0.2~\\msol \\simlt \\mcent \\simlt 1~\\msol$; see Fig. $6a$). During this period, $ 6~ \\msol~{\\rm{Myr}}^{-1} \\simlt \\mdot \\simlt 9~ \\msol~{\\rm{Myr}}^{-1}$ for $r \\simgt 500~\\rm{AU}$ (see Fig. $6g$). [Note, however, that the temperatures of HL Tauri and L1551-IRS5 are in the range $15 - 50~{\\rm K}$, which is greater than our assumed value of 10 K and should lead to higher values of $\\mdot$ and $\\mcent(t)$; e.g., Shu et al. 1987. For a discussion of how quantities scale with temperature in our models, see Basu \\& Mouschovias 1994.] In Figure 7 we show the mass ($M - \\mcent$) of the gas surrounding the central point mass in our typical model as a function of $r/\\rzero$ for the same seven times $\\delt_j$ as in Figure 6. (Taken together, Figs. $6a$ and 7 may be taken to represent the evolution of a protostar from a Class 0 to a Class I object.) For times $\\simgt \\delt_5$, the surrounding disk mass spans the range $0.01 - 0.1~\\msol$ for $r \\simgt 500~\\rm{AU}$, which agrees with the ${}^{13}{\\rm{CO}}$ disk masses of HL Tauri and L1551-IRS5 cited above. Finally, we note that the age of the oldest part of the molecular outflow from L1551-IRS5 is estimated to be $\\sim 10^5~\\rm{yr}$ (Bachiller, Tafalla, \\& Cernicharo 1994). This age is consistent with the time $\\delt$ needed for the central mass in our typical model to become $\\simgt 0.3~\\msol$ (see Fig. $2e$). Another observational consequence of our model is the magnetic field structure in the core after PMF. Figures $6b$ and $6d$ show that $\\BrZ \\approx \\Beq$ inside the core for the radius range $ 2 \\times 10^{-4} \\simlt r/\\rzero \\simlt 4 \\times 10^{-3}$. Hence, there is significant curvature of field lines (though, as discussed in \\S~3.3, there is less bending than there would be if the field had remained frozen into the neutrals), with bending angles $\\theta_B \\approx \\arctan(\\BrZ/\\Beq)$ in the range $20^{\\circ}- 50^{\\circ}$ for $180~ {\\rm AU} \\simlt r \\simlt 3.5 \\times 10^3~{\\rm{AU}}$. \\footnote{In agreement with the results of CM94 and CM95, we find that the magnetic tension force is generally not negligible in comparison with the magnetic pressure force at any radius (in either the core or the envelope), both before and after PMF.} This type of field geometry may be described roughly as having an hourglass shape. In fact, sub-mm polarimetry of the cores of W3 IRS5 (Greaves, Murray, \\& Holland 1994), Mon R2 (Greaves, Holland, \\& Murray 1995), and OMC-1 (Schleuning 1998) find field geometries suggestive of an hourglass shape on sub-parsec scales. In contrast, field lines remain essentially straight and parallel in the magnetically supported envelope ($r > 0.1~\\rm{pc}$), even after PMF. This agrees with polarimetric observations of molecular clouds that indicate well-ordered fields on these scales (e.g., Hildebrand, Dragovan, \\& Novak 1984; Hildebrand 1989, 1996; Novak et al. 1989; Kane et al. 1993; Hildebrand et al. 1995). A unique prediction of our model is the large ion--neutral drift speed that occurs during the post-PMF epoch. As shown in Figure $6f$, effective ambipolar diffusion following PMF yields $\\vd \\approx |\\vn| \\gg C$ in the inner regions of the core. In our typical model we find $\\vd \\simgt 1~\\rm{km}~{\\rm s}^{-1}$ for $\\delt \\simgt 2 \\times 10^4~\\rm{yr}$ on scales $r \\simlt 2 \\times 10^{-4} \\rzero \\simeq 180~\\rm{AU}$. Large drift speeds between neutrals and ions (such as $\\HCOp$, $\\rm{HCN}^+$, $\\rm{DCO}^+$, to name but a few) on these scales are therefore expected in our model. Detection of such drift speeds (through high-resolution observations of HCN or $\\HCOp$, say) could be used to observationally confirm our model results and to distinguish them from those of nonmagnetic collapse calculations (e.g., Shu 1977, Hunter 1977, Foster \\& Chevalier 1993) or of magnetic collapse models that do not account for the effect of ambipolar diffusion (e.g., Tomisaka 1996; Li \\& Shu 1997). \\subsection{Features of the Hydromagnetic Shock} The properties of hydromagnetic shocks in partially ionized gases have been developed extensively by many other authors (e.g., Mullan 1971; Draine 1980; Chernoff 1987; Roberge \\& Draine 1990; Draine \\& McKee 1993; Smith \\& Mac Low 1997). Because the ion {\\Alf} speed $v_{\\rm{A,i}} = \\Beq/(4 \\pi \\mi \\nni)^{1/2} = (\\mn/\\mi)^{1/2} (\\nn/\\nni)^{1/2} \\van$ is much larger than $\\van$, $|\\vn|$, and $|\\vi|$ in our model, we expect the outward-propagating shock that develops after PMF to have a magnetic precursor. This is indeed what we find in our simulation: the jump in the ion speed $\\vi$ and the magnetic field strength $\\Beq$ typically occurs at a distance of 1 to 3 computational mesh spacings further away from the symmetry axis than the jump in the neutral infall speed $\\vn$ and the column density $\\sign$. The displacement between the locations of the head of the disturbance in the neutral fluid and in the plasma and magnetic field decreases at later times. Examination of Figures $5b$, $5c$, and $6e$ reveals that, in the reference frame of the shock, the preshock infall speeds are supersonic, and the postshock speeds are also supersonic or just slightly subsonic. Similarly, the ion infall speeds (in the frame of the shock) are much less than the ion {\\Alf} speed. Therefore the shock we observe in our model is probably best classified as being C-type. \\footnote{A C-type shock is characterized by neutral velocities that (in the shock frame) remain supersonic throughout. Hence the shock in our simulation cannot be strictly of this type when the downstream neutral speed is subsonic. In that case a viscous (J-type) subshock may form (Draine \\& McKee 1993), although, as noted by Li \\& McKee (1996), turbulent diffusivity behind a real shock could plausibly keep the postshock flow supersonic and thereby obviate the need for such a subshock.} Making the approximation that in the vicinity of the shock the predominant magnetic stress is that due to the magnetic pressure gradient, the ion force equation becomes \\begin{equation} \\label{ionforceq} \\frac{\\sign}{\\tni} \\vd = - \\frac{Z}{4 \\pi} \\frac{\\partial \\Beq^2}{\\partial r} \\end{equation} (see eqs. [28c] and [51] in CM93, which contain additional terms, involving in particular the magnetic tension force, that could be used to refine the following simple estimate). This yields an approximate shock width \\leteq \\begin{eqnarray} \\label{shkapproxa} \\delshk &\\approx& \\frac{\\Bequ^2}{4 \\pi \\sign} \\frac{\\tni Z}{\\vd} \\left[\\left(\\frac{\\Beqd}{\\Bequ}\\right)^2 -1 \\right] = C \\tni \\frac{\\left(\\vanu/C \\right)^2}{2 \\left(\\vd/C\\right)} \\left[\\left(\\frac{\\Beqd}{\\Bequ}\\right)^2 -1 \\right] \\\\ \\label{shkapproxb} &=& 7.8 \\times 10^{13} \\frac{\\left(T/10~{\\rm K}\\right)^{1/2}\\left(\\vanu/C\\right)^2}{\\left(\\mn/2.33~{\\rm{a.m.u.}}\\right)^{1/2}\\left(\\nni/0.1~\\cc\\right)\\left(\\vd/C\\right)} \\left[1 + 0.067 \\frac{\\left(\\mHII/2~{\\rm{a.m.u.}}\\right)}{\\left(\\mi/30~{\\rm{a.m.u.}}\\right)}\\right] \\nonumber \\\\ &&\\hspace{5em}\\times \\left[\\left(\\frac{\\Beqd}{\\Bequ}\\right)^2-1\\right]~\\rm{cm}\\ , \\end{eqnarray} \\beq where $\\Bequ$ and $\\Beqd$ are the values of $\\Beq$ upstream and downstream of the shock, and $\\vanu$ is the upstream {\\Alf} speed. In deriving the last equality of equation (\\ref{shkapproxa}) we have used the relation $\\sign= 2 \\rhon Z$; equation (\\ref{tnieq}) and the relation $C= \\left(\\kB T/\\mn\\right)^{1/2}$ have been used in deriving equation (\\ref{shkapproxb}). At the time $\\delt_6$ the shock front is located at $r \\simeq 3.9 \\times 10^{-3} \\rzero = 5.2 \\times 10^{16}~{\\rm cm}$, and, in the vicinity of the front, $\\nni \\simeq 10^{-2}~\\cc$, $\\vanu \\simeq C$, $\\Beqd/\\Bequ \\simeq 4.6$ (see Fig. $6b$), and $\\vd \\simeq 0.4 C$ (see Fig. $6f$). For these values our rough estimate for the shock width given by equation (\\ref{shkapproxb}) yields $\\delshk \\simeq 4.3 \\times 10^{16}~{\\rm{cm}}$. Examination of Figure $6e$ at the time $\\delt_6$ reveals that the shock has an actual width $\\delshk \\simeq 2.1 \\times 10^{-3} \\rzero = 2.8 \\times 10^{16}~\\rm{cm}$. (Thus $\\delshk/r_{\\rm shk} \\approx 0.5$ at that time, so the thin-shock approximation that underlies the estimate [17] is marginally satisfied.) As noted in \\S~1, Li \\& McKee (1996) proposed that a hydromagnetic shock would form in a collapsing core as a result of the decoupling of flux from the ion and neutral fluids because of Ohmic dissipation (a process that becomes important in regions of density $\\nn \\gg 10^{11}~\\cc$; see \\S~2.1). They argued that the accumulating flux diffuses outward to regions of lower density, where improved coupling with the matter causes it to present an obstacle to the infalling neutral gas --- thereby giving rise to a hydromagnetic shock. While our numerical results have confirmed Li \\& McKee's basic shock-formation scenario, they have revealed that ambipolar diffusion, which mediates the shock, can also, following PMF, halt the inward advection of magnetic flux on scales ($r \\simgt 5~\\rm{AU}$) where Ohmic dissipation is not important. In other words, our results have shown that, during the post-PMF epoch, the field--matter decoupling that drives the hydromagnetic shock is due to ambipolar diffusion alone and does not depend on the effect of Ohmic dissipation at $r < \\rinner$. Because of the similarity in the basic shock-formation mechanism, it is of interest to compare our detailed simulation results with the predictions of the (simplified and analytic) shock model of Li \\& McKee (1996). From their requirement that the magnetic pressure of the shock balance the ram pressure of the neutrals, which were assumed to be freely-falling into the shock, Li \\& McKee derived relations for the shocked magnetic field strength and the shock location (see their eqs. [7] and [8]) in terms of the accretion rate $\\mdot$, the flux-to-mass ratio in units of the critical value for collapse (dubbed $\\epsilon$ in their paper), a parameter related to the logarithmic gradient of the magnetic field (dubbed $\\chi$), the ratio $Z/r$ of the local gas scale height and the radius (dubbed $h$), and the protostellar mass (denoted by $m_{\\ast}$). At the time $\\delt_6$ we have at the location of our shock $\\dot{M} \\approx 9~\\msol~{\\rm{Myr}}^{-1}$, $\\epsilon \\approx 0.9$, $\\chi \\approx 2$, $h \\approx 0.3$, and $m_{\\ast}=\\mcent(\\delt_6) \\approx 1~\\msol$. Inserting these values into their equations (7) and (8) yields a shocked magnetic field strength $\\sim 630~\\mu{\\rm G}$ and a shock radius $\\sim 2.1 \\times 10^3 ~{\\rm{AU}}$; by comparison, in our model the shocked magnetic field strength at that time is $\\Beqd \\approx 330~\\mu\\rm{G}$ and the shock radius is $r_{\\rm shk} \\approx 3.5 \\times 10^3~\\rm{AU}$ (see Figs. $6b$ and $6e$). The main reason why the analytic expression overestimates the numerically calculated magnetic field strength is that, contrary to the assumption of Li \\& McKee, the preshock neutrals are not in free fall but, rather, are strongly decelerated by magnetic forces (in our simulation we find that the preshock acceleration of the neutrals is reduced to $\\simeq 0.25 \\gr$). The corresponding reduction in the preshock ram pressure leads to a lower value of the postshock field amplitude, with a further reduction in the calculated field strength brought about by the contribution of magnetic tension (ignored in the analytic estimate) to the total magnetic force. Since the analytic estimate of $r_{\\rm shk}$ is based on relating the postshock field strength to the total magnetic flux inside the shock, the overestimate of the field strength naturally results in an underestimate of the shock radius. Despite these discrepancies, Li \\& McKee's analytic representation of the shock parameters provides a decent approximation to the results of our numerical calculation. Comparison of the column density $\\sign$ (see Fig. $6c$) and neutral infall speed $\\vn$ (see Fig. $6e$) behind the shock shows qualitative agreement with Figures $2a$ and $2d$ of Li \\& McKee (1996), including the free-fall behavior near the central protostar. In the pre-shock region, however, our infall speeds are smaller than theirs because of their assumption of free fall upstream of the shock: Li \\& McKee typically overestimate $\\vnu$ by a factor $\\sim 2$. As a result, the Mach number of the shock relative to the upstream gas (and thus the shock strength) is greater in their model than in ours (see Fig. $5c$) by a similar factor. We have not compared our results for the magnetic field structure behind the shock with those of Li \\& McKee on account of the fact that their system of MHD equations was not closed (it did not include the magnetic induction equation), so that they were unable to calculate the magnetic field with any accuracy (see their Fig. $2b$). \\subsection{Stability of the Core Against Magnetic Interchange} Our simulation has revealed that rapid ambipolar diffusion occurs behind the outward-propagating HMD. The effect that this has on the mass in the flux tubes downstream of the HMD can be seen in Figure $8a$, which shows the local mass-to-flux ratio $d M/d \\phiB = \\sign/\\Beq$ (normalized to the critical value for collapse) as a function of $r/\\rzero$ for the same seven times $\\delt_j$ as in Figure 6. For $\\delt > \\delt_1$ a local minimum in $\\sign/\\Beq$ appears after the passage of the HMD. Hence, there is a region behind the HMD for which $d (\\sign/\\Beq)/dr > 0$. This is a necessary condition for the onset of an interchange instability (e.g., Spruit \\& Taam 1990; Lubow \\& Spruit 1995; Spruit, Stehle, \\& Papaloizou 1995). Li \\& McKee (1996) speculated that such a situation could arise in the wake of a hydromagnetic shock in a collapsing core and suggested that it would act as source of turbulence in the postshock region of the inflow. \\footnote{Li \\& McKee (1996) also noted that the shock may be unstable to the Wardle instability, which involves ions collecting in magnetic field ``valleys'' (Wardle 1990). However, the shock will be immune to this instability if the ion density is determined by the local chemical reaction balance (as assumed in our calculation) rather than by the divergence of the ion mass flux.} Blaes \\& Balbus (1994) found that the magnetic shearing instability in differentially rotating disks could be stabilized if the disk is weakly ionized. This will also be the case for the interchange instability in a weakly ionized disk: instability is possible only if the growth rate $\\gaminst$ and the neutral--ion collision time $\\tni$ satisfy the condition $\\gaminst \\tni < 1$. This condition reflects the fact that there has to be sufficient collisional coupling between the ion and neutral fluids for a magnetic interchange instability to grow in the neutrals; otherwise the instability is damped. Calculation of $\\gaminst$ in our model is hampered by the fact that previous studies of interchange instability have been carried out only for disks that are in hydrostatic equilibrium, with exact balance between magnetic and gravitational forces. We can apply the results of these studies to our model only if the region behind the shock where $d (\\sign/\\Beq)/dr > 0$ is effectively in quasi equilibrium, with approximate balance between gravitational and magnetic forces, and with infall speeds $|\\vn| \\ll (r |\\gr|)^{1/2}$ ($\\simeq$ the free-fall speed). In our model, the magnitude of the acceleration of the neutrals in this region of the core does not become $\\simlt 0.1 | \\gr |$ until times $\\sim \\delt_6$; hence, approximate equilibrium between magnetic and gravitational forces is valid only at these later times. Spruit \\& Taam (1990) found that the growth rate for the most unstable linear interchange modes is \\begin{equation} \\label{growtheq} \\gaminst = \\left( \\frac{\\Beq \\BrZ}{2 \\pi \\sign} \\frac{d}{dr} \\ln \\frac{\\sign}{\\Beq} \\right)^{1/2} . \\end{equation} The product $\\gaminst \\tni$ for the region of the core susceptible to interchange instability is shown in Figure $8b$ as a function of $r/\\rzero$ at the time $\\delt_6$. We also plot (Fig. $8c$) the product $\\gaminst \\tau_{\\rm{kin}}$, where $\\tau_{\\rm{kin}} \\equiv r/|\\vn|$ is the kinematical timescale, as a function of $r/\\rzero$ for the same region of the core and time. If this product is $ < 1$, the unstable modes will be ``swept up'' by the infalling gas before they have time to grow. It is evident from these figures that $\\gaminst \\tni < 1$ and $\\gaminst \\tau_{\\rm{kin}} > 1$ for the region of the core susceptible to magnetic interchange. This means that there is sufficient collisional coupling between the ions and neutrals, and that the instability will grow before being swept along with the neutrals. Hence, this region of the core may be interchange unstable. An instability of this type would enhance the tansfer of gas with a high mass-to-flux ratio to the center (e.g., Spruit \\& Taam 1990), and, as noted by Li \\& McKee (1996), might also lead to the development of turbulence that could increase the field diffusivity in the postshock gas. However, the onset and development of this instability can only be studied by means of a fully 3-D simulation. \\subsection{Implications to the Magnetic Flux Problem} The magnetic flux problem in star formation has to do with the fact that the magnetic flux of a $1\\msol$ blob of matter in the diffuse interstellar medium is typically several orders of magnitude greater than the flux of a $1\\msol$ protostar. Such a blob of matter would therefore have to get rid of most of its flux before becoming a star. Ambipolar diffusion has long been suggested as a means by which the magnetic flux problem could be resolved (e.g., Mestel \\& Spitzer 1956; Mouschovias 1978; Paleologou \\& Mouschovias 1983; Nakano 1984; Mouschovias, Paleologou, \\& Fiedler 1985). In general, these earlier studies focused primarily on the role of ambipolar diffusion and the magnetic flux problem for the pre-PMF phase of protostellar evolution. \\footnote{On the basis of a consideration of the timescales for ambipolar diffusion and Ohmic dissipation at high densities, Nakano \\& Umebayashi (1986b) suggested that significant flux loss could only take place (primarily by Ohmic dissipation, according to their estimates) during the dynamical phase of core collapse. Lizano \\& Shu (1989) similarly concluded that the resolution of the protostellar magnetic flux problem must occur during the dynamical stage of core evolution: using the quasi-static approximation (valid for $\\nnc \\simlt \\rm{a~few} \\times 10^4~\\cc$; Fiedler \\& Mouschovias 1993; CM94; Basu \\& Mouschovias 1994) to calculate the contraction of a slightly subcritical molecular cloud, they found that only a small amount of flux is lost by ambipolar diffusion from the central flux tubes before runaway collapse is initiated.} While ambipolar diffusion does indeed reduce the flux-to-mass ratio during that phase, the flux contained within a $1\\msol$ region of a molecular cloud core is still much larger at the time of PMF than typical protostellar fluxes. Specifically, CM94 and CM95 found that the central $1\\msol$ flux tube within their cores had a total magnetic flux $\\sim 10^{30}~\\rm{Mx}$ (consistent with our results in \\S~3.2) during the pre-PMF dynamical collapse phase of the typical model. This value represents a reduction by a factor $\\sim 5.6$ of the flux associated with that mass before the onset of ambipolar diffusion. Nevertheless, it greatly exceeds the plausible upper limit ($\\sim 6 \\times 10^{26}~\\rm{Mx}$) on the flux of a solar-mass protostar (estimated assuming an average surface field of 10 kG and a stellar radius of $10^{11}~{\\rm cm}$; see Li \\& McKee 1996). We have shown in this paper that the rate of ambipolar diffusion is strongly increased during the post-PMF epoch of star formation. It is therefore of interest to examine the implications of our simulation results to the magnetic flux problem. As discussed in \\S~2.2, in our calculations we only consider the core regions at radii $r \\simgt \\rinner$ ($\\simeq 7.3~\\rm{AU}$ for our typical model), where ambipolar diffusion is the dominant mechanism of flux loss. Initially, the magnetic flux contained within $\\rinner$ is $\\phicent(\\delt=0) = 6.7 \\times 10^{26}~\\rm{Mx}$. As shown in Figure $3a$, the central flux increases before the onset of rapid ambipolar diffusion. This continues to the time $\\delt \\approx 10^3~\\rm{yr}$. For $\\delt > 10^3~\\rm{yr}$, ambipolar diffusion prevents further advection of flux from $r > \\rinner$ into the central sink, and $\\phicent$ changes very little after this time. By the time $\\delt \\approx 10^5~\\rm{yr}$, when $\\mcent \\approx 1\\msol$, $\\phicent \\approx 5 \\times 10^{27}~\\rm{Mx}$. This represents a decrease of over two orders of magnitude relative to the flux associated with this mass at the time of PMF. While this value is still about an order of magnitude higher than our adopted upper limit on the protostelar flux, the discrepancy is now much lower than previous estimates of ambipolar diffusion have indicated. The important conclusion from our work is thus that {\\em ambipolar diffusion in contracting molecular cloud cores can in principle contribute significantly to the resolution of the magnetic flux problem} by reducing the magnetic flux brought into a solar-mass protostar by a factor $\\simgt 10^3$. The new, and somewhat surprising, result is that {\\em most of this reduction occurs after PMF.} The remainder of the protostellar magnetic flux could possibly be extracted from the infalling mass through Ohmic dissipation within $\\rinner$, although refreezing of the magnetic field into the matter, brought about by collisional reionization at densities $\\nn \\simgt 10^{14}~\\cc$ (e.g., Pneuman \\& Mitchell 1965; Nakano \\& Umebayashi 1986b; Li \\& McKee 1996), as well as anomalous diffusivity (e.g., Norman \\& Heyvaerts 1985) operating in the reionized gas, could complicate the issue. Another complicating factor is the strong likelihood that much of the mass and flux carried into the protostar pass through a rotationally supported, circumstellar accretion disk of size $\\gg \\rinner$ (e.g., Lubow, Papaloizou, \\& Pringle 1994; Reyes-Ruiz \\& Stepinski 1996; Li 1996). It is also conceivable that magnetic flux is brought to the vicinity of the protostar but excluded from its interior by turbulent diffusivity associated with convection. Since the region within $\\rinner$ was excluded from our calculation, we do not pursue this topic any further in this paper." }, "9804/hep-ex9804007_arXiv.txt": { "abstract": "With an effective telescope area of order $10^4$~m$^2$, a threshold of $\\sim$50~GeV and a pointing accuracy of 2.5~degrees, the AMANDA detector represents the first of a new generation of high energy neutrino telescopes, reaching a scale envisaged over 25 years ago. We describe its performance, focussing on the capability to detect halo dark matter particles via their annihilation into neutrinos. ", "introduction": "\\unskip High energy neutrino telescopes are multi-purpose instruments; their science mission covers particle physics, astronomy and astrophysics, cosmology and cosmic ray physics. Their deployment creates new opportunities for glaciology and oceanography, possibly geology of the earth's core\\cite{pr}. Astronomy with neutrinos does have definite advantages. They can reach us, essentially without attenuation in flux, from the largest red-shifts. The sky is, in contrast, partially opaque to high energy photons and protons because of energy-loss suffered in interactions with infrared light, CMBR photons and radio waves\\cite{cronin}. They do not reach us from distances much larger than tens of megaparsecs once their energy exceeds thresholds of 10~TeV for photons and $5\\times 10^7$~TeV for protons. (Below this energy charged protons do not point back to their sources.) The drawback is that neutrinos are difficult to detect: the small interaction cross sections that enable them to travel without attenuation over a Hubble radius, are also the reason why kilometer-scale detectors are required in order to capture them in sufficient numbers to do astronomy\\cite{halzenkm}. Some opportunities may, however, be unique. If, for instance, the sources of the highest energy cosmic rays are beyond $10^2$~Mpc, conventional astronomy is unlikely to discover them. Some science missions do not require a detector of kilometer size. The best opportunities to search for halo dark matter are, in fact, associated with the present instrument which, while smaller in telescope area than the planned extension of AMANDA to ICE3 (ICECUBE), has a lower threshold. At this meeting, the capability of neutrino telescopes to discover the particles that constitute the dominant, cold component of the dark matter is of special interest. The existence of the weakly interacting massive particles (WIMPs) is inferred from observation of their annihilation products. Cold dark matter particles annihilate into neutrinos; {\\it massive} ones will annihilate into {\\it high-energy} neutrinos which can be detected in high-energy neutrino telescopes. This so-called indirect detection is greatly facilitated by the fact that the earth and the sun represent dense, nearby sources of accumulated cold dark matter particles. Galactic WIMPs, scattering off nuclei in the sun, lose energy. They may fall below escape velocity and be gravitationally trapped. Trapped WIMPs eventually come to equilibrium and accumulate near the center of the sun. While the WIMP density builds up, their annihilation rate into lighter particles increases until equilibrium is achieved where the annihilation rate equals half of the capture rate. The sun has thus become a reservoir of WIMPs which we expect to annihilate mostly into heavy quarks and, for the heavier WIMPs, into weak bosons. The leptonic decays of the heavy quark and weak boson annihilation products turn the sun and earth into nearby sources of high-energy neutrinos with energies in the GeV to TeV range. Figure~1 displays the neutrino flux from the center of the earth calculated in the context of supersymmetric dark matter theories\\cite{scopel}. The direct capture rate of the WIMPs in germanium detectors is shown for comparison. Contours indicate the parameter space favored by grand unified theories. Most of this parameter space can be covered by improving the capabilities of existing detectors by 2 orders of magnitude. Existing neutrino detectors have already excluded fluxes of neutrinos from the earth's center of order 1~event per $1000 \\rm~m^2$ per year. The best limits have been obtained by the Baksan experiment\\cite{suvorova}. They are already excluding relevant parameter space of supersymmetric models. We will show that, with data already on tape, the AMANDA detector will have an unmatched discovery reach for WIMP masses in excess of 100~GeV. \\begin{figure}[t] \\centering \\hspace{0in}\\epsfxsize=4.5in\\epsffile{fig1.eps} \\caption{Direct and indirect detection rates (for neutrinos from the center of the earth in the figure shown) of cold dark matter particles predicted by supersymmetric theory. Grand unified theories favor the parameter space indicated. Part of it is already excluded by present experiments as indicated by the horizontal and vertical lines.} \\end{figure} The potential of neutrino telescopes as dark matter detectors has been documented in detail\\cite{kamionkowski}. With a sensitivity which increases with the WIMP mass, they are complementary to direct, cryogenic detectors. They can detect WIMPS beyond the kinematic limits of the LHC: about 500~GeV for neutralinos. A striking way to illustrate their potential is to use the possible detection\\cite{belli} in the DAMA NaI detector in the Gran Sasso tunnel as an example. If their seasonal variation is indeed evidence for WIMPS, observation of a signal in an exposure of 4500 kg\\,days requires a WIMP-nucleon cross section of $10^{-42}{\\sim}10^{-41}$~cm$^2$ for a WIMP mass of $50{\\sim} 150$~GeV. This information is sufficient to calculate their trapping and annihilation rate in the sun and earth. Both will be a source of, on average, 100 neutrinos per year of WIMP origin in the existing AMANDA detector with an effective area of $10^4$~m$^2$. The exact rate varies with the mass of the WIMP. ", "conclusions": "" }, "9804/astro-ph9804107_arXiv.txt": { "abstract": "We have computed line profiles from self-gravitating toroids around black holes. The specific angular momentum of the toroids is assumed to be constant in space. The images of the toroids show peculiar feature in the rear side of the black holes. Concerning the line profiles, the red wing extends to the very small frequency region because the location of the inner edge is rather near the event horizon of the black hole and consequently the velocity of the inner edge of toroids can be faster than that of the Kepler disks. ", "introduction": "Recently, strong evidence for existence of supermassive black holes in the central regions of galaxies has been found from observations of electromagnetic waves with various wavelengths ranging from the radio waves to X-rays (see e.g. Miyoshi et al. 1995 for the radio observation; Ford et al. 1994, Harms et al. 1994 for the optical observation; Tanaka et al. 1995 for the X-ray observation). Radio and optical observations have shown that there exist very rapidly rotating gaseous disks in the central regions of galaxies. Since the high velocity of a disk implies the existence of a large amount of mass within a region of a very small size, it is widely considered that there are supermassive black holes with masses of $10^7 \\sim 10^9 M_{\\sun}$ at the centers of galaxies. However, such observations do not reveal the nature of black holes because the size of the observed region is still too large to get detailed information about the black holes. Contrary to these optical and radio analyses, recent X-ray observations have brought us important information about gravitational fields very near the massive objects or black holes. By using the ASCA satellite, broad iron emission lines have been detected in active galaxies (see e.g. Fabian et al. 1994; Mushotzky et al. 1995). In particular, Tanaka et al.~(1995) observed the Seyfert 1 galaxy, MCG--6--30--15, and discovered a very broad and skewed iron emission line. The broadness and skewness of the line profile can be explained only by assuming that the inner edge of the accretion disk is located very near the event horizon of the central black hole (Tanaka et al. 1995; Fabian et al. 1995). It implies that X-ray observations can be used to understand the very vicinity of the black holes. It is important to know the gravitational field near the event horizon because it is the key feature to distinguish a Schwarzschild black hole from a Kerr black hole. Consequently we have reached a stage to investigate very strong gravitational fields of black holes and to be able to test the validity of general relativity. However, since the emission line profiles depend not only on the gravitational fields but also on the structures of the accretion disks, it is not easy to determine the type of the black hole, i.e. whether the central black hole is of a Schwarzschild type or of a Kerr type. In fact, concerning the nature of the black hole at the center of MCG--6--30--15, various discussions have not settled down to a unique interpretation yet (see e.g. Tanaka et al. 1995; Iwasawa et al. 1996; Dabrowski et al. 1997; Bromley et al. 1997; Reynolds \\& Begelman 1997; Bromley et al. 1998). Therefore, in order to get a consistent picture of black hole -- accretion disk systems, we have to pursue much more investigations theoretically as well as observationally. Concerning the theory about the spectra of accretion disks, Cunningham~(1975) was the first to formulate the problem for Kerr black holes and obtained theoretical spectra from accretion disks around black holes (see also Cunningham 1976). Gerbal \\& Pelat~(1981) investigated lines from a ring around a black hole and found double-peaked asymmetric profiles. From the end of 80's, observations of lines in the X-ray spectrum stimulated many authors to study line profiles of accretion disks as well as disk structures (e.g. Nandra et al. 1989; Fabian et al. 1989; Kunieda et al. 1990; Kojima 1991; Laor 1991; Chen \\& Halpern 1989, 1990). However, in almost all theoretical studies mentioned above, investigations of the emission line profiles have been done by assuming that disks are geometrically thin and that only direct photons are observed. Some authors have studied the effect of multiple images (Luminet 1978; Bao, Hadrava \\& {\\O}stgaard 1994) and that of self-eclipse due to toroidal configurations (Bao \\& Stuchlik 1992; Kojima \\& Fukue 1992). For toroidal configurations the rotation law of the toroid is not always that of the Kepler rotation because of the presence of the pressure within the toroid. In fact, Kojima \\& Fukue~(1992) employed a variety of rotation laws. However, their analysis was done in the framework of Newtonian gravity. Therefore, quantitative treatment of non-Keplerian toroids in general relativity has not been carried out yet. Furthermore, in some situations, self-gravity of disks or toroids plays an important role for the structures of disks or toroids. In particular, massive neutron toroids around neutron stars or black holes have been proposed as possible sources of $\\gamma$-ray bursts (e.g. Paczy\\'nski 1991; Narayan, Paczy\\'nski \\& Piran 1992; Jaroszy\\'nski 1993; Witt et al. 1994). Although there might be little chance to observe line profiles from such systems even if exist, it would be interesting to study the effect of self-gravity of toroids on the line profiles. Such self-gravitating disks were investigated by Karas, Lanza \\& Vokrouhlicky~(1995). However, they have studied very light thin disks whose mass is less than several percent of the mass of the black hole. From the theoretical point of view, it would be interesting to investigate more massive disks or toroids as well as the effect of geometrical thickness. By treating massive toroids, the central objects are no more of Schwarzschild type nor Kerr type black holes because of the gravitational effect of the self-gravitating massive toroids on the black holes (Nishida \\& Eriguchi 1994). Moreover, gravitational effect of massive toroids may bring some differences to line profiles. Therefore, in this paper, we will consider self-gravitating toroid -- black hole systems and study their effect on emission line profiles. ", "conclusions": "In this paper we have constructed toroid -- black hole systems in general relativity. We have computed photon trajectories in the numerically obtained gravitational fields and investigated the images of the toroids as well as the line profiles from the toroids. In these computations, we have assumed that the observer is located very near the black hole because the metric in the whole space has not been calculated. Therefore, the quantitative values in this paper will be changed a little if the observer is located at infinity. However, the important thing is not to obtain exact quantitative values but to know characteristic feature which appears only by introducing thickness and self-gravity numerically exactly. In this sense, our results serve as representative ones for self-gravitating toroids around black holes. As discussed in Introduction, the observation of MCG--6--30--15 has given us time varying line profiles whose interpretation has not been clarified yet (Tanaka et al. 1995; Iwasawa et al. 1996; Dabrowski et al. 1997). Some authors have proposed models for the system which would explain the observations (Dabrowski et al. 1997; Bromley et al. 1997; Bromley et al. 1998). However, since there are many parameters about the disk structures and the X-ray sources, it is very difficult to obtain a unique solution for the system. This can be also seen from our result. For our Model B with the emissivity indices $s = -1$ and $s = -4$, the line profiles are shown in Fig.~\\ref{comparison} together with the observational data. As seen from this figure, the tendency of the observational data seems to be roughly explained by the change of the emissivity index of Model B. Thus, at the present stage, we can only say that since the model cannot be uniquely determined, we have to get more accurate observational data to clarify the spacetime of the central region of galaxies. \\begin{figure*} \\begin{tabular}{cc} (a)&(b)\\\\ \\epsfile{file=tanaka.eps,width=0.5\\textwidth} &\\epsfile{file=iwasawa.eps,width=0.5\\textwidth}\\\\ \\end{tabular} \\caption{ Same as Fig. 4 but for the model with $\\theta_0 = 30 \\degr$ and emissivity indices (a) $s=-1$ (solid line) and (b) $s=-4$ (solid line). The observational data of the ASCA are plotted by points with error bars for the data of Tanaka et al.~(1995) (a) and that of Iwasawa et al.~(1996) (b). } \\label{comparison} \\end{figure*}" }, "9804/astro-ph9804331_arXiv.txt": { "abstract": "We report the detection of the GRB 971214 counterpart in the near-infrared by means of two images in the K$^{\\prime}$-band taken at Calar Alto only $\\sim$3.5 and $\\sim$5 hours after the gamma-ray event. We detect the transient at K$^{\\prime}$=$18.03\\pm0.18$ and K$^{\\prime}$=$18.00\\pm0.22$ respectively. Our data seem to indicate the existence of a plateau with duration $ 1.5 \\leq T \\le 6.7 $ hours (between 3.5 and 10.2 hours after the high-energy event). Moreover the power-law decline should be steeper than the one given by the index $\\alpha_{\\rm K^{\\prime}}=0.45$. There is also a change in the slope of the broad-band spectrum at some wavelength between the J and K$^{\\prime}$ bands (possibly around the H-band). ", "introduction": "After 31 years since the discovery of gamma-ray bursts (GRBs), the origin of such brief gamma-ray flashes remains unknown. The observed isotropy of GRBs in the sky, could only be explained by theoretical models where GRBs originate either in a extended halo around the Galaxy or arise from sources at cosmological distances. Before the launch of the BeppoSAX satellite, the poor localization capability of the GRB detectors made the searches at other wavelengths unfruitful. The breakthrough took place in 1997 when the first X-ray afterglows were observed by the BeppoSAX, RXTE, ROSAT and ASCA satellites (Costa et al. 1997, Heise et al. 1997a, Marshall et al. 1997, Greiner et al. 1998, Murakami 1998). They were able to localize the fading X-ray emission that followed the more energetic gamma-ray photons once the GRB event had ended. This emission (the afterglow) extends to longer wavelengths, and the good accuracy in the position determination by BeppoSAX (typically $1^{\\prime}$ radius error boxes) has led to the discovery of the first optical counterparts for GRB 970228 (van Paradijs et al. 1997, Guarnieri et al. 1997), and GRB 970508 (Bond 1997, Djorgovski et al. 1997, Castro-Tirado et al. 1998), greatly improving our understanding of these puzzling sources. The measurement of the redshift for the GRB 970508 optical counterpart (Metzger et al. 1997) has established that one GRB, maybe all, lie at cosmological distances. GRB 971214 is the third GRB with a known optical counterpart. It was detected by the BeppoSAX Gamma-ray Burst Monitor (GBM, Frontera et al. 1997) on Dec 14.97 1997, as a 25 s long-structured gamma-ray burst. Simultaneous to the detection of the GBM, the Wide Field Cameras (WFC, Jager et al. 1997) on board BeppoSAX provided an accurate position (a $3^{\\prime}.9$ radius error box at a 3$\\sigma$ confidence level, Heise et al. 1997b) that allowed deep optical, infrared and radio observations. The position was also consistent with the one given by the all-sky monitor on RXTE (Bradt et al. 1993) and by the BATSE/Ulysses annulus (Kippen et al. 1997). When BeppoSAX pointed its Narrow-Field Instruments (NFI) to the GRB position, on Dec 15.25 ($\\sim$ 6.5 hours after the burst), a previously unknown variable X-ray source was found inside the WFC error box (Antonelli et al. 1997) which was identified as the X-ray afterglow of GRB 971214. Soon after, Halpern et al. (1997) reported the presence of a fading object inside the WFC GRB error box, based on two I-band images separated 24 hours. The object was afterwards confirmed as the counterpart of GRB 971214 by means of additional observations at other wavelengths: R-band (Castander et al. 1997, Diercks et al. 1998), I-band (Rhoads 1997) and J-band (Tanvir et al. 1997). No detections in the K-band were reported in the literature, although observations performed on Dec 15.54 imposed an upper limit of K $>$ 18.5 (Garcia et al. 1997). As it will be explained later, this upper limit will be used to constraint the power-law index $\\alpha_{\\rm K^{\\prime}}$ and the position of the possible maximum of the light curve. We report here the detection of the GRB 971214 counterpart in the near infrared (IR) by means of two K$^{\\prime}$-band images taken at Calar Alto on Dec 15.12 and 15.18 (mean observing time, only $\\sim$3.5 and $\\sim$5 hours after the gamma-ray event). The second image is almost simultaneous to the beginning of the observations performed by the BeppoSAX narrow-field instruments. We discuss whether our observations are in agreement with the extrapolation of the power-law seen at other bands in later epochs. ", "conclusions": "We have detected the GRB 971214 near-IR counterpart $\\sim$3.5 hours and $\\sim$5 hours after the gamma-ray event which enables to conclude that: i) a magnitude difference $\\Delta$K${^{\\prime}}=-0.03\\pm 0.28$ is derived from our measurements, whereas $\\Delta$K${^{\\prime}}=0.464$ would be expected assuming a power-law decay with index $\\alpha_{\\rm K^{\\prime}}=1.2$ (similar to the one observed at optical wavelengths). This implies a deviation of $1.7\\sigma$. If the assumed power-law index $\\alpha_{\\rm K^{\\prime}}$ were 1.4, then the rejection level would be $2.0\\sigma$. Thus, our measurements suggest a rising or a flat light curve segment with a duration $ 1.5 \\leq T \\le 6.7 $ hours (between 3.5 and 10.2 hours after the burst). This conclusion must be taken with care since the above-mentioned rejection levels are not stringent enough to assure the result with total confidence; ii) the power-law decline in the near-IR should be steeper than the one given by $\\alpha_{\\rm K^{\\prime}}=0.45$; iii) for the observations carried out on Dec 15.44-15.51, there is a change in the slope of the measured energy distribution at some wavelength between the J and K$^{\\prime}$ bands (possibly around H)." }, "9804/astro-ph9804041_arXiv.txt": { "abstract": "The measurements of CMB anisotropy have opened up a window for probing the global topology of the universe on length scales comparable to and beyond the Hubble radius. For compact topologies, the two main effects on the CMB are: (1) the breaking of statistical isotropy in characteristic patterns determined by the photon geodesic structure of the manifold and (2) an infrared cutoff in the power spectrum of perturbations imposed by the finite spatial extent. We present a completely general scheme using the {\\em regularized method of images} for calculating CMB anisotropy in models with nontrivial topology, and apply it to the computationally challenging compact hyperbolic topologies. This new technique eliminates the need for the difficult task of spatial eigenmode decomposition on these spaces. We estimate a Bayesian probability for a selection of models by confronting the theoretical pixel-pixel temperature correlation function with the {\\sc cobe--dmr} data. Our results demonstrate that strong constraints on compactness arise: if the universe is small compared to the `horizon' size, correlations appear in the maps that are irreconcilable with the observations. If the universe is of comparable size, the likelihood function is very dependent upon orientation of the manifold {\\it wrt} the sky. While most orientations may be strongly ruled out, it sometimes happens that for a specific orientation the predicted correlation patterns are preferred over the conventional infinite models. ", "introduction": "The remarkable degree of isotropy of the cosmic microwave background (CMB) points to homogeneous and isotropic Freidmann-Robertson-Walker (FRW) models for the universe. The underlying Einstein's equations of gravitation are purely local, completely unaffected by the global topological structure of space-time. In fact, in the absence of spatially inhomogeneous perturbations, a FRW model predicts an isotropic CMB regardless of the global topology. The observed large scale structure in the universe implies spatially inhomogeneous primordial perturbations exist which gave rise to the observed anisotropy of the CMB. The global topology of the universe does affect the local observable properties of the CMB anisotropy. In compact universe models, the finite spatial size usually precludes the existence of primordial fluctuations with wavelengths above a characteristic scale related to the size of the universe. As a result, the power in the CMB anisotropy is suppressed on large angular scales. Another consequence is the breaking of statistical isotropy in characteristic patterns determined by the photon geodesic structure of the manifold. One can search for such patterns statistically in the COBE maps, and to the extent that they are not there, one can constrain the size of the universe and its topology. Much recent astrophysical data suggest the cosmological matter density parameter, $\\Omega_0$, is subcritical~\\cite{opencase}. If a (possibly varying) cosmological constant is absent or insufficient to bring the total density to the critical value, this would imply a hyperbolic spatial geometry for the universe (commonly referred to as the `open' universe in the cosmological literature). The topologically trivial (simply connected) hyperbolic 3-space, $\\hm$, is non-compact and has infinite size. There are numerous theoretical motivations, however, to favor a spatially compact universe~\\cite{motive}. To reconcile a compact universe with a flat or hyperbolic geometry, consideration of spaces with non trivial topology (non simply connected spaces) is required. A compact cosmological model can be constructed by identifying points on the standard infinite flat or hyperbolic FRW spaces by the action of a suitable discrete subgroup, $\\Gamma$, of the full isometry group, $G$, of the FRW space. The infinite FRW spatial hypersurface is the {\\em universal cover}, tiled by copies of the compact space (most appropriately represented as the {\\em Dirichlet domain} with the observer at its {\\em basepoint}). Any point ${{\\bf x}}$ of the compact space has an image ${{\\bf x}}_i = \\gamma_i {{\\bf x}}$ in each copy of the Dirichlet domain on the universal cover, where $\\gamma_i \\in \\Gamma$. The hyperbolic manifold, $\\hm$, can be viewed as a hyperbolic section embedded in four dimensional flat Lorentzian space. The isometry group of $\\hm$ is the group of rotations in the four space -- the proper Lorentz group, $SO(3,1)$. A compact hyperbolic (CH) manifold is then completely described by a discrete subgroup, $\\Gamma$, of the proper Lorentz group, $SO(3,1)$. The Geometry Centre at the University of Minnesota has a large census of CH manifolds and public domain software SnapPea~\\cite{Minn}. We have adapted this software to tile $\\hm$ under a given topology using a set of generators of $\\Gamma$. The tiling routine uses the generator product method and ensures that all distinct tiles within a specified tiling radius are obtained. A CH manifold, ${\\cal M}$, is characterized by a dimensionless number, ${\\cal V}_{\\!\\cal M}\\equiv V_{\\!\\cal M}/d_c^3$, where $V_{\\!\\cal M}$ is the volume of the space and $d_c$ is the curvature radius \\cite{Thur7984}. There are a countably infinite number of CH manifolds with no upper bound on ${\\cal V}_{\\!\\cal M}$. The smallest CH manifold discovered so far has ${\\cal V}_{\\!\\cal M} =0.94$~\\cite{smallestCH}. ~\\footnote{ The volume of CH manifolds is bounded from below and the present theoretical lower bound stands at ${\\cal V}_{\\!\\cal M} \\ge 0.167$~\\cite{gab_mey96}. There exist sharper lower bounds within subclasses of CH manifolds under restrictions on topological invariants~\\cite{minvol}. It has been conjectured that the smallest known manifold is in fact the smallest possible~\\cite{smallestCH}.} The Minnesota census lists several thousands of these manifolds with ${\\cal V}_{\\!\\cal M}$ up to $\\sim 7$. In the cosmological context, the physical size of the curvature radius $d_c$ is determined by the density parameter and the Hubble constant $H_0$: $d_c=(c/H_0)/\\sqrt{1-\\Omega_0}$. The physical volume of the CH manifold with a given topology, \\ie a fixed value of $V_{\\!\\cal M}/d_c^3$, is smaller for smaller values of $\\Omega_0$. Two quantities which characterize linear dimensions of the Dirichlet domain are $R_>$ and $R_<$, the radii of circumscribing and inscribing spheres, respectively. In the standard picture, the CMB that we observe is a Planckian distribution of relic photons which decoupled from matter at a redshift $\\approx 1100$. These photons have freely propagated over a distance $ R_{\\sc ls} \\approx 2 d_c\\, {\\rm arctanh} \\sqrt{1-\\Omega_0} $, comparable to the ``horizon'' size. For the adiabatic fluctuations we consider here, the dominant contribution to the anisotropy in the CMB temperature measured with wide-angle beams ($\\theta_{\\sc fwhm} \\gta 2^\\circ \\Omega_0^{1/2}$) comes from the cosmological metric perturbations through the Sachs-Wolfe effect. The adiabatic cosmological metric perturbations can be expressed in terms of a scalar gravitational potential, $\\Phi({\\bf x},\\tau)$. The dynamical equation for the gravitational potential allows for separation of the spatial and temporal dependence,\\footnote{At the scales appropriate to CMB anisotropies, damping effects on $\\Phi$ can be neglected.} $\\Phi({\\bf x},\\tau) = F(\\tau) \\Phi({\\bf x},\\tauls)$, where $F(\\tau)$ encodes the time dependence of the metric perturbations and $\\Phi({\\bf x},\\tauls)$ is the field configuration on the three-hypersurface of constant time $\\tau=\\tauls$ when the last scattering of CMB photons took place. We shall study open, $\\Omega_0<1$, models with zero cosmological constant where in the matter dominated phase,~\\cite{mukh92} \\begin{equation} F(\\tau) = \\frac{5(\\sinh^2\\tau -3 \\tau\\sinh\\tau +4\\cosh\\tau -4)} {(\\cosh\\tau -1)^3}\\,. \\lbl{Feta} \\end{equation} Here and further on we use dimensionless conformal time $\\tau$ expressed in units of the curvature radius. A non-zero cosmological constant can be trivially incorporated in our analysis by using the appropriate solution for $F(\\tau)$. We write the Sachs-Wolfe formula for the CMB temperature fluctuation, $\\Delta T(\\hat q)$, in a direction $\\hat q$, in the form \\begin{equation} \\fl \\dT(\\hat q) = \\frac{1}{3} \\Phi(\\hat q\\chiH,\\tauls) + 2 \\int_{0}^\\chiH d\\chi f(\\chi) \\Phi(\\hat q\\chi,\\tauls)\\,, \\quad f(\\chi)=\\frac{d}{d\\tau} F(\\tau)\\bigg|_{\\tau=\\chiH-\\chi}\\,, \\lbl{dTSW} \\end{equation} where $\\chi$ is the affine parameter along the photon path from $\\chi=0$ at the observer position to $\\chiH=R_{\\sc ls}/d_c$. The first term is called the {\\em surface} or ``naive'' Sachs-Wolfe effect (NSW). The second term, which is nonzero only if $\\Phi$ varies with time between $\\tauls$ and now, is the {\\em integrated} Sachs-Wolfe effect (ISW). The angular correlation between the CMB temperature fluctuations in two directions in the sky is then given by \\begin{eqnarray} \\fl C(\\hat q,\\hat q^\\prime)\\equiv \\left\\langle\\dT(\\hat q)\\dT(\\hat q^\\prime)\\right\\rangle = \\frac{1}{9} \\langle\\Phi(\\hat q\\chiH,\\tauls)\\Phi(\\hat q^\\prime\\chiH,\\tauls)\\rangle \\nonumber\\\\ \\lo +\\frac{2}{3}\\int_{0}^\\chiH d\\chi~f(\\chi)~\\left[ \\langle\\Phi(\\hat q\\chi,\\tauls)\\Phi(\\hat q^\\prime\\chiH,\\tauls)\\rangle +\\langle\\Phi(\\hat q^\\prime\\chi,\\tauls)\\Phi(\\hat q\\chiH,\\tauls)\\rangle\\right] \\nonumber\\\\ \\lo +4\\int_{0}^\\chiH d\\chix{1} ~f(\\chix{1})~ \\int_{0}^\\chiH d\\chix{2}~f(\\chix{2})~ \\langle\\Phi(\\hat q\\chix{1},\\tauls)\\Phi(\\hat q^\\prime\\chix{2},\\tauls)\\rangle\\,. \\lbl{cthetaSW} \\end{eqnarray} The main point to be noted is that $C(\\hat q,\\hat q^\\prime)$ depends on the spatial two point correlation function, $\\xi_\\Phi \\equiv \\langle\\Phi({\\bf x},\\tauls)\\Phi({\\bf x^\\prime},\\tauls)\\rangle $ of $\\Phi$ on the three-hypersurface of last scattering. This is due to the fact that the equation of motion for $\\Phi$ allows a separation of spatial and temporal dependence. Although in this work we restrict our attention to the Sachs-Wolfe effect which dominates when the beam size is large, we should point out that other effects which contribute to the CMB anisotropy at finer resolution can also be approximated in terms of spatial correlation of quantities defined on the hypersurface of last scattering~\\cite{us_inprep}. ", "conclusions": "" }, "9804/astro-ph9804088_arXiv.txt": { "abstract": "We have imaged the inner square arcminute of the well known lensing and cooling flow cluster A2390 ($z = 0.23$) down to a sensitivity of 65 and 130 $\\mu$Jy at 6.75 and 15 $\\mu$m , respectively. We report the first evidence of an active star-forming region in a cooling flow (at those wavelengths) and strong emission in the mid-IR from lensed galaxies located at $z=0.9$. ", "introduction": "The cluster of galaxies A2390 (z=0.228) possesses remarkable properties which makes its study particularly attractive: presence of a ``straight'' giant gravitational arc (z=0.913), numerous arclets, an elongated galaxy distribution (\\cite{Mellier} ; \\cite{Pello}) and a large velocity dispersion (1093 km s$^{-1}$, Carlberg et al., 1996) as well as a high X-ray luminosity ($\\sim1.5 ~10^{45}$ erg s$^{-1}$ in the [0.1--2.4] keV band). A deep HRI ROSAT pointing revealed an elongated X-ray morphology, the existence of a secondary maximum responsible for the observed gravitational shear in the optical and a strong cooling flow of $\\sim800$ M$_{\\odot}$yr$^{-1}$ (Pierre et al., 1996). All this indicates that A2390, and its underlying gravitational potential, is especially relevant for our understanding of massive cluster formation, which is, in a hierarchical scenario, closely related to the history of galaxy/star formation. This has motivated deep ISOCAM observations of the cluster core during the guaranteed time programme DEEPXSRC. We present here the observations and results of the cD galaxy and the $z=0.9$ lensed system. Throughout the paper we assume $H_{o} = 50$ km s$^{-1}$Mpc$^{-1}$ and $q_{o}$ = 0.5. ", "conclusions": "In this letter, we restrict the discussion to the 4 sources found in the maximum sensitivity area of ISOCAM rasters and seen both at 6.7 $\\mu$m and 15 $\\mu$m, i.e. \\#1--4. Their photometric properties are summarized in Table \\ref{sources}. The visible and near-IR spectral energy distributions (SED) computed for these 4 sources are given in Table \\ref{SED}, and compared to those obtained for a typical cluster galaxy. Details on these photometric data can be found in Pell\\'o et al. (1998) and the references therein. In addition, about 20 objects are identified at 6.7 $\\mu$m and 10 at 15 $\\mu$m, which will be discussed in a forthcoming paper. All sources are point-like for ISOCAM, except the cD galaxy which extends over two times the PSF FWHM at 6.7 $\\mu$m (i.e $\\sim 20$ kpc), and then allow us to exclude a pure AGN emission. \\subsection{The cD galaxy and its cooling flow} The cD galaxy is detected both at 6.7 $\\mu$m and 15 $\\mu$m, with a flux of $300^{+50}_{-40} \\mu$Jy and $500^{+80}_{-70} \\mu$Jy respectively. VLA observations (\\cite{Arnouts}) show a point-like source with decreasing radio fluxes of 140, 16 and 5.5 mJy at 6, 2 and 1.3 cm, respectively. Assuming a power law spectrum, the mid-IR flux would be some $10^5$ fainter than observed, which excludes a jet-like synchrotron contribution to the observed mid-IR emission. In galaxies where the mid-IR emission is dominated by an old stellar population, the ratio 6.7 $\\mu$m/15 $\\mu$m is $> 1$. An excess of 15 $\\mu$m emission in field galaxies indicates the presence of dust heated by UV photons from star-forming regions. Compared to the 6.7 $\\mu$m/15 $\\mu$m ratios observed in other nearby early-type galaxies (\\cite{Suzanne}) or in distant cD galaxies in clusters (\\cite{next}), the ratio of $\\sim 0.6$ found here for the cD is exceptional. This ratio is compatible with the colors of the disk component of Centaurus A (\\cite{Felix}), a nearby giant early-type galaxy exhibiting active star-forming regions in dust lanes, due to a merge with a spiral galaxy. Thus, the cD in A2390 is probably also undergoing active star formation. However, our cD galaxy looks notably different in other wavelengths. Cen A shows a jet plus an extended radio emission but our cD does not. In addition, a B image reveals the existence of a filament extended along the main axis of the cD, while V and I HST images show the presence, within the filament, of very blue globules possibly associated with the 15 $\\mu$m maximum (Fig. \\ref{zoom}). Strong emission lines are present across the long-slit spectrum of the filament (Fig. \\ref{Halpha}) with ratios indicative of massive star formation associated with shocks and incompatible with an active nucleus (\\cite{Allen}, \\cite{Baldwin}). Finally the SED of the filament exhibits a clear excess in the V and B bands with respect to what is expected for a typical elliptical. Assuming that the V flux is mainly produced by forming stars, we derive $M_V = -20.8 \\pm 0.2$ and a SFR of $8 \\pm 4$ h$_{50}^{-2}$ M$_{\\odot}$ yr$^{-1}$ for the optical filament, in agreement with the values obtained from the B flux, corrected for the $[OII]3727$ emission. No absorption has been considered in this calculation, so this estimated SFR has to be taken as a lower limit. According to the results derived from V and B band, and from IR, the optical light is probably coming from the most external regions of the star-forming system, whereas part of the star-formation activity remains shrouded within more dense and dusty clouds, as in the Antennae Galaxies, where absorption is ten times higher when derived from mid-IR than from J, H, K bands, as most massives stars are not visible at optical wavelengths (\\cite{antenne}). This implies that we can not exclude a SFR as high as ten times what is derived from the optical for the cD of A2390. Those differences explain why, with a 6.7 $\\mu$m/15 $\\mu$m ratio of 0.6, we can not just consider the cD of A2390 as an early-type galaxy undergoing simple star-formation in dust lanes as Cen A, but most probably as the place of one or several massive star-bursts which may be located in the central globules (see Fig \\ref{zoom}). Indeed, the study of the X-ray image of this cluster demonstrated the presence of one of the strongest cooling flows known ($\\sim 800$ M$_{\\odot}$ yr$^{-1}$ within a cooling radius of 200 kpc), surrounding the cD galaxy (\\cite{marguerite}). Giving the size of the mid-IR emitting region in the cD, $\\sim 20$ kpc, we derive a mass flow of 80 M$_{\\odot}$ yr$^{-1}$, assuming that $\\dot{M} \\sim r$ (\\cite{fabian}). Note that 20 kpc is also about the size of the optical filament (Fig. \\ref{zoom}). However, our present understanding of the relationship between mid-IR dust emission and star formation is still too preliminary to infer quantitative constraints on the IMF or even on the heating processes involved in this complex medium. Finally, the total star-formation rate of $\\sim 10$ M$_{\\odot}$yr$^{-1}$ deduced from the optical in the filament is clearly a lower limit, and the huge quantity of gas needed could be provided by the cooling flow. However, despite the fact that spiral galaxies are very rare in the core of rich clusters, the hypothesis of a past merge with a late-type galaxy cannot be formally excluded here, which would also provide gas for some $10^{7-8}$ years. \\begin{figure} \\psfig{file=lemonon.center3.ps,width=8.5cm} \\caption []{The V-I color HST image, obtained from F555W and F814W, of the core of the cD galaxy of A2390. White and black crosses indicate 6.7 $\\mu$m and 15 $\\mu$m maximum emission respectively with the associated uncertainties (see text). Some very blue globules suspected to be active star forming regions could be responsible for the 15 $\\mu$m emission which coincides with the most luminous one, and associated with strong cooling flow ($\\sim800$ M$_{\\odot}$yr$^{-1}$) surrounding this cD galaxy. They are part of a NW--SE filament, aligned with the cD main axis, and the overall cluster X-ray elongation on large scale.} \\label{zoom} \\end{figure} \\begin{figure} \\psfig{file=lemonon.halpha.ps,width=9.8cm} \\caption []{Mean long-slit spectrum of the filament within the cD galaxy (without reddening correction), from CFHT (\\cite{Arnouts}).} \\label{Halpha} \\end{figure} \\subsection{What is new in the arc system of Abell 2390 ?} After the detection of the giant arc at $z=0.724$ in Abell 370 (\\cite{Metcalfe}), observation of the complex arc system of Abell 2390 confirms the capability of ISOCAM to point up very distant lensed objects. The giant arc consists of three parts, A at $z=1.033$ (\\cite{brenda}), and B--C, at $z=0.913$ (\\cite{Pello}). Near IR imaging already distinguished A from B--C, as A was not detected in the K band (\\cite{Smail}). HST images revealed that B and C are likely two interacting galaxies. The present ISOCAM images are in full agreement with this picture. Although it was not possible to estimate properly the 6.7 $\\mu$m flux because of blending, the 15 $\\mu$m/6.7 $\\mu$m ratio for the B--C component is well larger than unity which is indicative of the presence of an active star forming region in agreement with the strong [OII] line detected in the optical spectrum (\\cite{Pello}). Except for its lower amplification factor, the case of object D is very similar. Its morphology in the HST images is complex with probable signs of interaction and low surface brightness extensions. The existence of starburts in the two interacting galaxies is then not a surprise. The optical and near-IR SEDs of objects B--C and D appear brighter in the near-IR and fainter in the blue bands compared to A. These SEDs can be fitted by different synthetic spectra at $z = 0.913$, using the GISSEL96 code (Bruzual \\& Charlot, 1998) to approximately constrain the parameters, and a single stellar population (instantaneous burst), an extinction curve of SMC type (Pr\\'evot et al. 1984), and assuming the Scalo IMF (1986). The best fits of the sources B--C and D are obtained with a rest-frame $A_{V} \\sim3$ in both cases, a stable result with respect to metallicity changes. The corrected magnitudes for objects B--C and D (lensing and absorption) are very similar ($M_B = -20.8$) , the total mass involved in the burst being $\\sim10^{10} M_{\\odot}$ in both cases. Despite uncertainties on burst age, a constant star-forming model gives similar results and a mean corrected SFR of $40$ to $50\\, M_{\\odot} h_{50}^{-2} yr^{-1}$. According to these results, the two lensed sources detected by ISOCAM at $z=0.913$ are strongly reddened star-forming galaxies. In the case of A, there is no need for a reddening correction to fit the SED. Finally, the ISOCAM source \\#4 detected in both channels may be associated with a very faint source in the HST image (I = 23.5), with a fuzzy shape. Its 15 $\\mu$m/6.7 $\\mu$m ratio is very high ($\\sim 3$). A photometric redshift of $z=0.4^{+0.2}_{-0.08}$ is proposed for this object by techniques described by Miralles \\& Pell\\`o (1998). Even if the results are much more uncertain in this case ($75\\%$ confidence), the best fit of the SED gives $A_{V} \\sim$3.5--4.2 in order to explain the high J and K$^\\prime$ emission compared to the optical bands. The corrected SFR is relatively low, $\\sim1 M_{\\odot} h_{50}^{-2} yr^{-1}$. Taking the photometric redshift into account, the SED of this object, with strong mid-IR emission with respect to its optical counterpart, is probably dominated by the so-called unidentified infrared band emitters, and its colors are similar to those of the post-starburst companion of M51 (Boulade et al., 1996). \\subsection{Summary and conclusion} From deep and high-resolution ISOCAM images of the core of Abell 2390 we discovered active star forming regions in the two most distant lensed galaxies ever seen in a cluster by ISO. This allowed us to support the scenario of two interacting galaxies at $z=0.913$ in the ``straight arc'' of A2390, as well as in the other galaxy observed at the same redshift. More interesting, we detect a very faint emission from the cD galaxy at 6.75 $\\mu$m, compared to other cluster dominant galaxies at similar redshift (\\cite{next}). But, for the first time, the strong 15 $\\mu$m/6.75 $\\mu$m emission ratio flags the presence of a large amount of warm dust in the cD, probably associated with a very active star forming region located within the envelope of the galaxy. Thus, our observation may further elucidate the fate of part of accumulating gas in the complex cooling flow radio core environment." }, "9804/astro-ph9804277_arXiv.txt": { "abstract": "Using an Eulerian perturbative calculation, we show that the distribution of relative pairwise velocities which arises from gravitational instability of Gaussian density fluctuations has asymmetric (skewed) exponential tails. The negative skewness is induced by the negative mean streaming velocity of pairs (the infall prevails over expansion), while the exponential tails arise because the relative pairwise velocity is a {\\it number}, not volume weighted statistic. The derived probability distribution is compared with N-body simulations and shown to provide a reasonable fit. ", "introduction": "\\label{sec-intro} Redshift surveys present a distorted picture of the world because peculiar motions displace galaxies from their true spatial positions. This phenomenon, which would make redshift surveys useless for intergalactic spaceship navigators, is extremely useful for cosmologists. It can serve as a probe of the dynamics of gravitational clustering and the cosmological mass density parameter, $\\Omega$ (\\cite{sar77}; \\cite{pee80}, hereafter LSS; \\cite{kai87}; \\cite{ham92}; \\cite{pee93}, hereafter PPC; \\cite{reg95}). A convenient statistical measure of the distortion effect is the galaxy two-point correlation function in redshift space. Under certain assumptions it can be expressed as a convolution of the true spatial correlation function, $\\xi(r)$, with the distribution of the relative line-of-sight velocities of pairs of galaxies, $p(w|r,\\theta)$ . Here $r$ and $w$ are respectively, the spatial separation and relative radial velocity of a pair of galaxies, while $\\theta$ is the angle between the separation vector ${\\bf r}$ and observer's line of sight (cf. LSS; \\cite{fis95}, hereafter F95). The purpose of this {\\it Letter} is to derive $p(w|r,\\theta)$, using weakly nonlinear gravitational instability theory. This distribution was measured from N-body simulations and estimated indirectly from redshift surveys. At $r \\simlt$ 1\\mpc, where\\footnote{We use the standard parametrisation for the Hubble constant, $H = 100 \\, h$ \\kms Mpc$^{-1}$.} the galaxies are strongly clustered ($\\xi \\simgt 20$), the observations are consistent with an exponential distribution (\\cite{pee76}; \\cite{dav83}; \\cite{fis94}, hereafter F94; \\cite{mar95}; \\cite{lsd97}). The fact that $p(w)$ at small separations differs strongly from its initial, Gaussian character, is not surprising: after all, the small-scale velocity field has been `processed' by strongly-nonlinear dynamics in clusters, and exponential distributions were recently derived from the \\cite{pre74} theory (\\cite{she96}, \\cite{dia96}). On larger scales, where the fluctuations have small amplitudes, one na{\\\"\\i}vely expects to see the `unprocessed' initial conditions. However, N-body experiments suggest that $p(w|r,\\theta)$ retains its exponential character even at separations $r \\simgt$ 10\\mpc, where $\\xi \\simlt 0.1$, despite the fact that the initial density and velocity fields in those experiments were drawn from a Gaussian distribution (\\cite{efs88}, hereafter EFWD; \\cite{zur94}, hereafter ZQSW; F94). At similar separations, an exponential $p(w|r,\\theta)$ has also been inferred from observations (F94; \\cite{lov96}). The simulations also show that the radial component of the distribution, $p(w|r,0^{\\circ})$ is significantly skewed, in particular at large separations (EFWD; ZQSW; F94). The physical origin of the skewness and exponential shape of $p(w)$ at large separations has until now remained unexplained. We provide the explanation below. ", "conclusions": "" }, "9804/astro-ph9804310_arXiv.txt": { "abstract": "We have examined the hypothesis that the majority of the diffuse EUV flux in the Coma cluster is due to inverse Compton scattering of low energy cosmic ray electrons ($0.16 < \\epsilon < 0.31$\\,GeV) against the 3$^{\\circ}$\\,K black-body background. We present data on the two-dimensional spatial distribution of the EUV flux and show that these data provide strong support for a non-thermal origin for the EUV flux. However, we show that this emission cannot be produced by an extrapolation to lower energies of the observed synchrotron radio emitting electrons and an additional component of low energy cosmic ray electrons is required. ", "introduction": "Diffuse EUV emission has been detected in five clusters of galaxies: Virgo (Lieu et al., 1996a), Coma (Lieu et al., 1996b), Abell 1795 (Mittaz et al., 1997), and Abell 2199 and Abell 4038 (Bowyer et al., 1997). These clusters were detected with a statistical significance varying from 8 to 50 standard deviations. The diameter of the diffuse emission in these clusters ranges from 20$\\arcmin$ to 40$\\arcmin$. Some diffuse EUV emission in clusters of galaxies would be expected from the well-studied X-ray cluster emission. However, in all cases examined to date, the EUV emission is far greater than the expected emission from the X-ray-emitting gas. Marginal signatures of this ``soft excess'' are sometimes present in the lowest energy resolution band of the ROSAT PSPC, where they produce less than a 20\\% enhancement over the emission expected from the X-ray cluster gas. In contrast, the excesses found with EUVE range from 70\\% to 600\\% above that expected from the X-ray gas. A variety of instrumental and Galactic interstellar medium absorption effects have been suggested as alternative explanations for the EUVE data, but these have all been found wanting (For a discussion of these issues, see Bowyer et al., 1997). In the original reports of diffuse EUV cluster emission, the data were interpreted in terms of additional thermal gas components in the clusters. In these analyses, the known X-ray emission from the hot cluster gas was first fitted to the EUV data, and the excess EUV emission was computed. This excess was then fitted by additional components of thermal gas. Because of the low energy of the EUV emission, much lower temperature thermal gases ($\\sim 10^6$\\,K) are required. The concept that additional components of lower temperature gas are present in these clusters has not received wide support, primarily because gas at these temperatures is near the peak of the radiative cooling curve and hence cools rapidly, requiring a substantial energy input to sustain the gas at these temperatures. In addition, it is difficult to understand how different components of gas at grossly different temperatures could retain their separate identity. For example, the Coma cluster has been shown to be formed by merging of distinct subunits, in both X-ray (White et al., 1993), and optical (Colless and Dunn, 1996) studies. This produces variations less than a factor of 3 in temperatures of the X-ray emitting gas (Honda et al. 1996). Deiss and Just (1996) have shown in a general analysis, and with a specific application to the Coma cluster, that turbulent mixing time scales are only a few $10^9$\\,years, which argues against the co-existence of major quantities of gas at two vastly different temperatures. However, Cen et al. (1995) have argued that a warm ($\\simeq 10^6$\\,K) thermal gas is widely distributed throughout the universe, as a direct product of the growth of structure leading, eventually, to clusters of galaxies. In their scenario, the energy required to sustain the warm gas is provided by gravitation. Hwang (1997) has examined the hypothesis that the source of the diffuse EUV flux is inverse Compton scattering by electrons that are a low energy extrapolation of electrons producing the observed synchrotron emission; these electrons are scattered against the 3$^{\\circ}$\\,K black-body background radiation. The magnetic field he derived for the cluster was 0.2 to 0.4\\,$\\mu$G which is consistent with the range of estimates for the cluster field. In this work he only considered the constraints imposed by the total EUV flux. En{\\ss}lin and Biermann (1998) also considered this mechanism as the source of the EUV flux in the Coma cluster. They assumed that the relativistic energy density of the synchrotron emitting electrons scale radially with the same profile as the X-ray producing gas. This assumption can be questioned since the non-thermal relativistic electrons may well be independent of the thermal X-ray gas. These authors cite the best support for this assumption is given by the data in Figure\\ 3 of Deiss et al. (1997) which compares the X-ray and radio radial emission profiles. Unfortunately, the results in this figure are incorrect as has been confirmed by Deiss (private communication). These authors find a magnetic field of 1.2\\,$\\mu$G is required given their assumptions; this field is also consistent with the range of estimates for the cluster. Sarazin and Lieu (1998) have explored the possibility that EUV radiation in clusters of galaxies could be produced by inverse Compton scattering by a population of very low energy cosmic ray electrons. They showed that the one dimensional EUV spatial profile for the cluster A1795, a radio quiet cluster, was consistent with this hypothesis. A potential problem with this hypothesis as a universal explanation for the EUV emission in clusters of galaxies is that the electrons they proposed have an energy density and pressure which are 1 to 10\\% of that of the thermal gas in clusters. If one includes the pressure of cosmic ray ions to the pressure of the electrons proposed by Sarazin and Lieu, using the ratio expected on theoretical grounds (Bell 1978) and measured at Earth orbit (see. e.g., Weber 1983), the total cosmic ray pressure is substantially larger than that of the X-ray emitting gas. In this work we reconsider the hypothesis that the EUV emission in the Coma cluster is the result of inverse Compton emission. We first consider the constrains imposed by the total EUV flux. We review the existing radio data and obtain a different spectral index than that employed by Hwang (1997) and En{\\ss}lin and Biermann (1998), which we argue is more appropriate. We then derive results which are generally consistent with the inverse Compton hypothesis. We then consider the two dimensional spatial distribution of the EUV flux; this data provides substantial support for a non-thermal origin for this flux. We find that the spatial distribution of the magnetic field required by En{\\ss}lin and Biermann to produce the EUV emission profile is unrealistic. We show that the difference in spatial extent between the EUV and radio halos cannot be explained using an electron distribution which is an extrapolation of the known synchrotron emitting electrons and that an additional population of low energy cosmic ray electrons are required to explain these data. ", "conclusions": "We have examined the hypothesis that the EUV radiation from the Coma cluster is due to inverse Compton scattering of low energy cosmic ray electrons against the 3$^{\\circ}$\\,K black-body background radiation. The total integrated EUV emission produced by cosmic ray electrons which are a low energy extrapolation of higher energy electrons, known to be present from their synchrotron emission, gives results which are consistent with the range of estimates of the magnetic field in the cluster. We next consider the two dimensional spatial distribution of the EUV emission. This emission does not follow the distribution of the gravitationally bound X-ray gas, but rather exhibits an asymmetric distribution similar to that exhibited by the radio emission. This suggests a non-thermal origin for the EUV emission rather than a gravitationally constrained thermal gas. We show from a comparison of the size of the EUV halo and the radio halo that the EUV emission cannot be produced by inverse Compton radiation from electrons which are an extrapolation of the distribution which produces the observed radio emission. We develop a model for the EUV emission which is self-consistent and fits the existing data. This model requires an additional component of low energy cosmic rays. Inverse Compton EUV emission is surely present at some level in clusters of galaxies with radio halos. However, it may well be masked by emission from some other more dominant source mechanism. A test of the inverse Compton scattering hypothesis as the source of the EUV flux in the Coma cluster would be provided by a measurement of the size of the radio halo at $\\sim 1$\\,MHz. Unfortunately, this is a challenging measurement because of instrumental limitations and ionospheric effects. In addition, at these low frequencies self-absorption could affect the surface brightness profile by reducing the flux near the cluster center while increasing the halo size; this effect would have to carefully be considered when interpreting such a measurement." }, "9804/astro-ph9804126_arXiv.txt": { "abstract": "Cosmological implications of clusters of galaxies are discussed with particular attention to their importance in probing the cosmological parameters. More specifically we compute the number counts of clusters of galaxies, \\ns relation, in X-ray and submm bands on the basis of the Press--Schechter theory. We pay particular attention to a set of theoretical models which well reproduce the {\\it ROSAT} 0.5-2 keV band \\ns, and explore possibilities to break the degeneracy among the viable cosmological models. ", "introduction": "There are several reasons why clusters of galaxies are regarded as useful probes of cosmology including (i) since dynamical time-scale of clusters is comparable to the age of the universe, they should retain the cosmological initial condition fairly faithfully. (ii) clusters can be observed in various bands including optical, X-ray, radio, mm and submm bands, and in fact recent and future big projects (e.g., SDSS, AXAF, PLANCK) aim to make extensive surveys and detailed imaging/spectroscopic observations of clusters. (iii) to the first order, clusters are well approximated as a system of dark matter, gas and galaxies, and thus theoretically well-defined and relatively well-understood, at least compared with galaxies themselves, and (iv) on average one can observe a higher-z universe with clusters than with galaxies. In particular X-ray observations are well-suited for the study of clusters since the X-ray emissivity is proportional to $n_e^2$ and thus less sensitive to the projection contamination which has been known to be a serious problem in their identifications with the optical data. In fact, various statistics related to the abundances of clusters has been extensively studied to constrain theories of structure formation, including mass function (Bahcall \\& Cen 1993; Ueda, Itoh, \\& Suto 1993), velocity function (Shimasaku 1993; Ueda, Shimasaku, Suginohara, \\& Suto 1994), X-ray Temperature function (hereafter XTF, Henry \\& Arnaud 1991; White, Efstathiou, \\& Frenk 1993; Kitayama \\& Suto 1996 ; Viana \\& Liddle 1996; Eke, Cole, \\& Frenk 1996; Pen 1996). Previous authors have focused on cosmological implications of cluster XTF mainly because theoretical predictions are relatively easier although the observational data are statistically limited. In addition, the conversion to the number density at high z becomes very sensitive to the adopted cosmological parameters. On the other hand, \\ns which we discuss in details below is observationally more robust (Ebeling et al. 1997; Rosati \\& Della Ceca 1997) while its theoretical prediction is more model-dependent (Oukbir, Bartlett, \\& Blanchard 1996; Kitayama \\& Suto 1997; Kitayama, Sasaki \\& Suto 1998). In this respect, both statistics are complementary. \\begin{figure}[t] \\begin{center} \\leavevmode \\psfig{figure=logns1.cps,height=6.8cm} \\psfig{figure=chi2ns2.cps,height=6.8cm} \\end{center} \\vspace*{-0.5cm} \\caption{ {\\it Left:} Theoretical predictions for \\ns of X-ray clusters in CDM models with different cosmological parameters; (a) $\\sigma_8=1.04$ models with different $\\Omega_0$, $\\lambda_0$ and $h$, (b) $\\Omega_0=1$ and $0.45$ models with different $\\sigma_8$. Denoted by (COBE) are the models normalized according to the {\\it COBE} 4 year data (Bunn \\& White 1997). Data points with error bars at $S\\simlt 10^{-12}$ \\unit are from the {\\it ROSAT} Deep Cluster Survey (RDCS, Rosati et al. 1995; Rosati \\& Della Ceca 1997), and the error box at $S\\simgt 2 \\times 10^{-12}$ represents a power-law fitted region from the {\\it ROSAT} Brightest Cluster Sample (BCS, Ebeling et al. 1997). For the BCS data at $S= 2 \\times 10^{-12}$, $1 \\times 10^{-11}$ and $6 \\times 10^{-11}$\\unit, we also plot the corresponding Poisson errors. {\\it Right:} Limits on $\\Omega_0$ and $\\sigma_8$ in CDM models ($n=1$, $h=0.7$) with (a) $\\lambda_0=1-\\Omega_0$, and (b) $\\lambda_0=0$. Constraints from cluster \\ns (solid) and XTF (dotted) are plotted as contours at $1 \\sigma$(68\\%), $2\\sigma$(95\\%) and $3\\sigma$(99.7\\%) confidence levels. Dashed lines indicate the {\\it COBE} 4 year results from Bunn \\& White (1997). } \\label{fig:ns1chi2} \\vspace*{-0.5cm} \\end{figure} ", "conclusions": "Let us summarize the conclusions of the present talk. \\begin{description} \\item{(1)} There exist { several theoretical models} which successfully reproduce the observed \\ns relation of galaxy clusters over almost four orders of magnitude in X-ray flux. \\item{(2)} The resulting $\\sigma_8$ is given by the following empirical fit (95\\% confidence limit): \\begin{equation} \\sigma_8 = (0.54 \\pm 0.02 \\pm 0.1) \\times \\Omega_0^{-0.35-0.80\\Omega_0+0.55\\Omega_0^2} \\end{equation} for $\\lambda_0=1-\\Omega_0$ CDM, and \\begin{equation} \\sigma_8 = (0.54 \\pm 0.02 \\pm 0.1) \\times \\Omega_0^{-0.28-0.91\\Omega_0+0.68\\Omega_0^2} \\end{equation} for $\\lambda_0=0$ CDM. \\item{(3)} Low-density CDM models ($n=1$) with $(\\Omega_0,\\lambda_0,h,\\sigma_8) = (0.3,0.7,0.7,1)$ and $(0.45, 0, 0.7, 0.8)$ simultaneously account for the cluster \\ns, XTF, the {\\it COBE} 4 year normalization. \\end{description} Maybe the most important point is that many cosmological models are more or less successful in reproducing the structure at redshift $z\\sim0$ {\\it by construction}. This is because the models have still several degrees of freedom or {\\it cosmological parameters} which can be appropriately {\\it adjusted} to the observations at $z\\sim0$ ($\\Omega_0$, $\\sigma_8$, $h$, $\\lambda_0$, $b(r,z)$). We have shown that, given a complete flux limited cluster sample with redshift and/or temperature information, one can further constrain the cosmological models. In fact, our tentative comparison indicates that our predictions reproduce well the evolutionary features of the XBACs and that the results, although preliminary, seem to favor low density ($\\Omega_0 \\sim 0.3$) universes. As indicated by this preliminary result, surveys of objects at at high redshifts in several different bands (X-ray, radio and submm) are the most efficient and rewarding to break the degeneracy among the viable cosmological models." }, "9804/astro-ph9804256_arXiv.txt": { "abstract": "Steep soft X-ray (0.1-2 keV) quasars share several unusual properties: narrow Balmer lines, strong FeII emission, large and fast X-ray variability, rather steep 2-10 keV spectrum. These intriguing objects have been suggested to be the analogs of Galactic black hole candidates in the high, soft state. We present here results from ASCA observations for two of these quasars: NAB0205+024 and PG1244+026. Both objects show similar variations (factor of $\\sim2$ in 10 ks), despite a factor of about ten difference in the 0.5-10 keV luminosity ($7.3\\times 10^{43}$ erg s$^{-1}$ for PG1244+026 and $6.4\\times10^{44}$ erg s$^{-1}$ for NAB0205+024, assuming isotropic emission, $H_0$ = 50.0 and $q_0$ = 0.0). The X-ray continuum of the two quasars flattens by 0.5-1 going from the 0.1-2 keV band toward higher energies, strengthening recent results on another half dozen steep soft X-ray AGN. PG1244+026 shows a significant feature in the `1 keV' region, which can be described by either as a broad emission line centered at 0.95 keV (quasar frame) or as edge or line absorption at 1.17 (1.22) keV. The line emission could be due to reflection from an highly ionized accretion disk, in line with the view that steep soft X-ray quasars are emitting close to the Eddington luminosity. Photoelectric edge absorption or resonant line absorption could be produced by gas outflowing at a large velocity (0.3-0.6 c). ", "introduction": "The ROSAT PSPC has found a large spread in the energy spectral indices of low-z quasars\\footnote{We use ``quasars'' to describe broad line emission objects, regardless of luminosity.} : $0.5<\\alpha_{0.1-2keV}<3.5$. In about 10$\\%$ of cases $\\alpha_{0.1-2keV}\\gs 2$ (e.g. Laor et al. 1994, 1997, Walter \\& Fink 1993, Fiore et al. 1994). The large spread in $\\alpha_{0.1-2keV}$ favoured the discovery of its correlation with other properties. In fact, the steep soft X-ray quasars have then been realized to share a cluster of unusual properties: \\begin{itemize} \\item narrow Balmer lines \\footnote{the permitted lines have FWHM$\\ls$2000~km~s$^{-1}$, yet still are clearly broader than the forbidden lines.} ( Laor et al 1994, 1997, Boller et al. 1995); \\item strong FeII emission (Laor et al 1994, 1997, Lawrence et al. 1997) \\item Rapid, large amplitude variability (factor of 2-50 on timescales from minutes to months, Boller et al., 1995, Brandt et al., 1995, Otani 1995, Boller et al. 1997) \\item Somewhat steep hard X-ray spectra (2$>\\alpha_{2-10keV} >0.6$, Pounds et al. 1995, Brandt et al., 1997); \\end{itemize} Pounds et al. (1995), suggest the latter to be a close physical analogy with the X-ray power-law produced by Comptonization in a hot accretion disk corona in Galactic black hole candidates (BHC) in their `soft-high' state. This is not the only analogy between BHC and steep X-ray spectrum quasars. Laor et al. (1994, 1997) explained the correlation with H$\\beta$ FWHM as due to the larger size of a virialized broad emission line region for an AGN in a high $L/L_{Edd}$ state. Ebisawa (1991) found that while the soft component of 6 BHC observed by Ginga is roughly stable on time scales of 1 day or less, the hard component exhibits large variations down to msec time scales. These timescales translates to $10^4$ years and 0.1 day for quasars, if they scale with the mass of the compact object. The soft component of BHC extends up to $\\sim 10$ keV in BHC in `soft-high' states, and it is often associated with optically thick emission from an accretion disk. If this is the case, the temperature should scale with the mass of the compact object as $M_{BH}^{-1/4}$, and the above energy translates to 0.1-0.4 keV for quasars. The rapid large amplitude variability shown by a few narrow line Seyfert 1 galaxies (NLSy1) at about 1 keV on timescales of hours to days (Otani 1995, Brandt et al. 1995, Boller et al. 1997) can then be analogous to the above BHC hard component flickering. A steep X-ray spectrum quasar with 10-100 times the luminosity of NLSy1s, should be larger and so should vary no more rapidly than several days. Instead Fiore et al. (1998a) find that steep X-ray spectrum PG quasars commonly vary by a factor 2 in 1~day. Variability seems therefore correlated with X-ray spectral slope and Balmer line width (and therefore possibly with the accretion rate) rather than with the luminosity. Evidence for spectral features in the `1 keV' region in many steep soft X-ray quasars is building up (Turner et al., 1991, Brandt et al., 1994, Otani et al., 1995, Comastri et al., 1995, Leighly et al., 1997, 1998a,b). Instead, `normal' Seyfert 1 galaxies (having broad Balmer lines and flatter soft X-ray spectra) usually have their strongest absorption features at lower energies (in the 0.6-0.9 keV `oxygen' band). An intriguing possibility is that the appearance of these features at different energies also depend on $L/L_{\\rm Edd}$. Detailed high energy X-ray spectra of luminous quasars with steep soft X-ray spectra are essential to understand the `narrow-broad line' phenomenon in AGN, in particular whether the peculiar X-ray properties depend on optical luminosity, optical-to-X-ray ratio ($\\alpha_{OX}$), or on their Eddington ratio. To this end we selected two bright quasars with $\\alpha_{0.1-2keV}>$2.0 (Fiore et al., 1994) at the extreme values of optical luminosity, both with low Galactic $N_H$ (Table 1) of $1.9\\times 10^{20}$ cm$^{-2}$ for PG~1244+026, and of $3.0\\times 10^{20}$ cm$^{-2}$ for NAB0205+024, Elvis et al., 1989) and observed them with ASCA. We report the results in this paper. ", "conclusions": "ASCA observations of two steep soft X-ray quasars have shown that: \\begin {enumerate} \\item The X-ray continuum of the two quasars flattens by $\\Delta\\alpha=0.5-1$ going toward high energies. Similar results were obtained by authors on some half dozen steep soft X-ray quasars. \\item PG1244+026 shows a significant feature in the `1 keV' region. Similar features were again reported in other steep soft X-ray quasars. The data are not good enough to discriminate between a broad emission line centered at 0.95 keV (quasar frame) or an absorption edge at 1.17 keV, or an absorption notch at 1.22 keV. Line emission could be due to reflection from an highly ionized accretion disk, in line with the view that steep soft X-ray quasars are emitting close to the Eddington luminosity. Photoelectric edge absorption or resonant line absorption could be produced by gas outflowing at a large velocity (0.3-0.6 c). In these absorption models significant cold (i.e. oxygen less ionized than OVI) absorption in excess of the Galactic is required. This would imply an increase by a factor 2-3 of the cold column with respect to a previous PSPC observation or a peculiar ionization structure. In neither the emission or absorption cases the SIS resolution is good enough to identify unambiguously the ions responsible for the feature. The high resolution and high throughput of the low energy gratings and spectrometers of AXAF and XMM are clearly needed to shed light on this puzzling case. \\item The two quasars show similar variability properties (flux variations up to a factor of 2 in 10 ks) despite a factor of ten difference in the X-ray observed luminosity. This agrees with the Fiore et al (1998a) finding that the variability properties of radio-quiet quasars are correlated with the shape of the X-ray spectrum, the width of the Balmer lines and so possibly with the accretion rate. \\end{enumerate} \\bigskip F.F acknowledges support from NASA grants NAG 5-2476 and NAG 5-3039, B.J.W. acknowledges support from ASC contract NAS8-39073." }, "9804/astro-ph9804060_arXiv.txt": { "abstract": "Since it has become possible to discovery planets orbiting nearby solar-type stars through very precise Doppler-shift measurements, the role of methods used to analyze such observations has grown significantly. The widely employed model-dependent approach based on the least-squares fit of the Keplerian motion to the radial-velocity variations can be, as we show, unsatisfactory. Thus, in this paper, we propose a new method that may be easily and successfully applied to the Doppler-shift measurements. This method allows us to analyze the data without assuming any specific model and yet to extract all significant features of the observations. This very simple idea, based on the subsequent subtraction of all harmonic components from the data, can be easily implemented. We show that our method can be used to analyze real 16 Cygni B Doppler-shift observations with a surprising but correct result which is substantially different from that based on the least-squares fit of a Keplerian orbit. Namely, using frequency analysis we show that with the current accuracy of this star's observations it is not possible to determine the value of the orbital eccentricity which is claimed to be as high as 0.6. ", "introduction": "Recent improvements in the long-term precision of Doppler-shift measurements \\cite[]{Marcy:96c::} resulted in several spectacular detections of planetary companions to solar-type stars \\cite[for review see the paper of][] {Marcy:97::}. As such discoveries supply indirect evidence of the existence of extra-solar planets, other explanations of observed radial-velocity variations appeared, e.g., stellar pulsations \\cite[]{Gray:97::}. The most recent results, however, show that only a planetary hypothesis is acceptable \\cite[]{Marcy:98::,Gray:98::}. The usual procedure showing that there exists a planet around s star consits of a direct least-square fit of Keplerian model to the observations. It always gives certain values of orbital parameters and their formal errors. In the case of `good' data this is the best and the quickest way to obtain relable results. However, in the case of spare data with big errors one has to prove that the least-squares method can be used and that the obtained parameter values and their errors are good estimates of the real values. This is a difficult and time consuming task. Without doubt, we have this situation with the Doppler observation of extra-solar planets. We present an analysis of this problem and we show that the eccentricity of the fitted orbit is a very sensitive parameter and, in some cases, its value and error given by the least-squares method are not correct. The aim of this paper is to show how the mentioned problem can be solved in practice. Namely, we propose a method that can be very useful for analyzing radial-velocity variations. It is based on a simple idea involving the subsequent subtraction of periodic components from the data. This approach allows us to analyze the observations without assuming any specific model describing the system behavior (like Keplerian motion or stellar pulsations). After the determination of all significant components of the data, it remains to be decided which process is responsible for what we observe and whether it is possible to choose only one. The plan of this paper is as follows. In Section 2 we analyse observations of 16 Cygni B and we explain why the standard least-squere fit does not give reliable estimates of parameters and their errors. In Section 3 we analyze analytically the Keplerian motion of the system `a star with one planet' in order to learn how its motion modulates the observed star radial-velocities. We investigate mainly the spectral properties of the motion which are essential for our method. In Section 4 we develop a simple numerical technique which can be used to extract all the information we need to compare with the results from Section 3. In Section 5, we perform a numerical test of the method using simulated radial-velocity variations with the orbital parameters of 70 Vir \\cite[]{Marcy:96b::}. In section 6, we discuss the application of the method to finding the eccentricity of 16 Cygni B \\cite[]{Cochran:97::}. ", "conclusions": "In this paper we have shown that results obtained from least-squares fits of Keplerian orbits to real Doppler-shift measurements may lead to incorrect interpretations. Specifically, they may give unrealistic or even entirely false values of parameters and their uncertainities. In order to solve these problems we have proposed a new method, frequency analysis, which efficiently provides an independent test of the reliability of determined orbital parameters. This method may deliver a substantial revision of the current values of planets' high eccentricities that are essential for our understanding of the formation and evolution of planetary systems. It might even lead to hints that some of the observed high eccentric planets are in fact planetary systems consisting of more than one planet or at least provide an independent point of view on the same data. These facts, together with the ease of applicability of frequency analysis, make our method worth trying on future observations if not for the data already gathered." }, "9804/physics9804027_arXiv.txt": { "abstract": "It is shown that the plasma, generated during an impact of a meteoroid with an artificial satellite, can produce electromagnetic radiation below the microwave frequency range. This interference is shown to exceed local noise sources and might disturb regular satellite operations. ", "introduction": "At the end of 1998, the first modules of the \\emph{International Space Station} will be put in orbit around the Earth and this should open new frontiers for life in space. The intensive use of the space makes necessary to know the potential risks. The threat from meteoroids is today well known and several authors have underlined the risks connected with the impact on a spacecraft (for a review, see \\cite{REV}). However, the \\emph{Olympus} end-of-life anomaly \\cite{CASWELL} and the recent work of McDonnell \\emph{et al.} \\cite{MCDONNELL} put a new light on these issues. The \\emph{Olympus} failure is a paradigmatic example: in that case, the impact with a Perseid meteoroid may have generated electrical failures, leading to a chain reaction which culminated with an early end of the mission \\cite{CASWELL}. On the other hand, McDonnell \\emph{et al.} \\cite{MCDONNELL} showed that, if the plasma charge and current production during an impact are considered, meteoroid streams can be very dangerous, even during normal conditions. It should be noted that they considered only damages by direct discharges or current injection in circuits (\\emph{e.g.} via the umbilical) \\cite{MCDONNELL}. However, there are several other ways by which the plasma could interact with the spacecraft electronics. For example, it is useful to recall the work of Cerroni and Martelli \\cite{CERRONI}, in which they showed that thermal forces in impact-produced plasmas could explain the magnetisation observed in the neighbourhood of lunar craters. Even if Cerroni and Martelli studied experimentally hypervelocity impacts of aluminium projectiles on basalt targets, it is possible to extend their work to general hypervelocity impacts. Here, we show that a plasma cloud, generated during a hypervelocity impact of a meteoroid with an artificial satellite, can radiate electromagnetic energy below the microwave frequency range and, therefore, may disturb regular satellite operations. ", "conclusions": "After the \\emph{Olympus} end-of-life anomaly \\cite{CASWELL} and the work of McDonnell \\emph{et al.} \\cite{MCDONNELL}, it seems clear that the meteoroids hazard is not restricted to a mechanical damage. Here it is suggested a new interference path, that is electromagnetic radiation emitted from the impact-produced plasma cloud. Even if the radiated power is not sufficient to destroy anything, it may disturb regular satellite operations. Further investigations should be made on specific satellite, because they require detailed information about onboard electronics, in order to calculate possible couplings and non-linearities. \\stars Author wishes to thank Paolo Farinella, of Department of Mathematics of the University of Pisa, for constructive review. \\newpage" }, "9804/astro-ph9804195_arXiv.txt": { "abstract": " ", "introduction": "The Hubble Deep Field (HDF) provides accurate multi-band photometry of galaxies to very faint magnitudes, with 10$\\sigma$ $AB$ magnitude limit of $m_{814}=27.60$ (\\cite{HDF}). The faint limit of the HDF makes it difficult to obtain spectroscopic redshifts for the majority of the galaxies in the field. It is therefore useful to derive estimated redshifts of these galaxies using the available multi-band photometry. Several groups (\\cite{Lanze96}, \\cite{GwHa96}, \\cite{Saw97}) obtained photometric redshifts for the HDF by comparing the observed $UBVI$ fluxes of each object with a set of galaxy spectral templates of different galaxy types redshifted to evenly spaced redshifts. Since spectroscopic redshifts have been measured and published for $\\sim$ 100 galaxies in the HDF (\\cite{Cohen96}, \\cite{Hogg98}; \\cite{Stei}; \\cite{DEEP}), it is possible to fit analytic expressions for photometric redshifts. In this paper, we explore a simple empirical approach to estimating redshifts of galaxies based on their colors (see \\cite{Conno97}, \\cite{Brun97}, \\cite{Conno95} for an alternative empirical approach); this method has the advantage of being simple, model independent (i.e., it does not depend on the assumption of any particular set of galaxy spectral templates), and easy to use in determining approximate redshifts of $z\\la 4$ galaxies. We determine the empirical analytic relations for color redshifts in \\S 2. We compare our estimated redshifts for the HDF galaxies with those obtained by the template-fitting method in \\S 3. We describe our estimated redshift catalog of HDF galaxies to $z\\la 4$ in \\S 4 (the Web site address of the catalog is given). We investigate the redshift clustering of HDF galaxies in \\S 5, and summarize our results in \\S 6. ", "conclusions": "Using HDF photometric and spectroscopic data, we have determined a set of simple analytic formulae that yield estimated galaxy redshifts to $z\\la 4$ in terms of linear combinations of three measured colors, $U-B$, $B-V$, and $V-I$ (Eqs.(\\ref{eq:Lz3,1})-(\\ref{eq:Hz2,2})). The derived analytic formulae in five color ranges exhibit small dispersions between the estimated and spectroscopic redshifts. For $z\\la 2$ galaxies, the redshift dispersion ranges from $\\sigma_z=0.034$ to $\\sigma_z=0.097$ for different color ranges. For $z\\ga 2$ galaxies, we find $\\sigma_z=0.14$ and $\\sigma_z=0.36$ for two color ranges which typically represent $z\\ga 3$ and $z\\la 3 $ galaxies respectively. These color-redshift relations apply to about 90% in the sample. The smallest dispersion between the color and the spectroscopic redshifts, $\\sigma_z=0.034$, occurs for the $z\\la 2$ galaxies satisfying $(U-B)< (B-V) -0.1$; 28 galaxies with measured redshifts are used in deriving the relation for the estimated redshift, with only 4 free parameters (the coefficients in Eq.(\\ref{eq:Lz3,1})). There are 230 HDF galaxies with $I< 27$ and measured UBVI magnitudes that belong to this color range; we investigate the large-scale redshift distribution of these galaxies and find evidence for peaks in the redshift distribution that suggest large-scale clustering to at least $z\\sim 1$. These results are consistent with those of Cohen et. al. (1996) using observed spectroscopic redshifts of a smaller number of galaxies. We have applied our color redshift formulae to the entire HDF photometric catalog and find that the derived redshifts are consistent with those obtained from spectral template-fitting techniques. The analytic relations, by design, yield lower dispersion than the template-fitting method. The color-redshift relations have the advantage of being simple, model independent, and easy to use. They can be further improved with additional data. These analytic color-redshift estimators are useful in providing empirical estimates of galaxy redshifts to $z\\la 4$ using multiband photometry. Our Estimated Redshift Catalog of HDF Galaxies, based on our color redshift formulae for all 848 HDF galaxies with $I<27$ and measured $UBVI$ fluxes, is available by anonymous ftp in the elt/:HDF subdirectory of astro.princeton.edu. Note that our color-redshift relations (Eqs.(\\ref{eq:Lz3,1})-(\\ref{eq:Hz2,2})) are derived using AB magnitudes and for the HDF filters. For application to other photometric catalogs, the appropriate spectroscopic training set should be used; when such a training set is not available, Eqs.(\\ref{eq:Lz3,1})-(\\ref{eq:Hz2,2}) may provide useful estimates after appropriate photometric transformation has been performed between the different filter systems. Also note that these color-redshift relations should not be applied to galaxies which lie close to the boundaries of the color ranges. Finally, we note that our color-redshift relations are limited by the absence of measured spectroscopic redshifts for galaxies in the range of $1.4 \\le z \\le 2.2$ (see Fig.1 and Fig.2). It is very important to obtain spectroscopic redshifts in this range, because it will not only enable better calibration of photometric redshifts, it will also help us understand the nature of galaxies in the intermediate redshift range." }, "9804/astro-ph9804148_arXiv.txt": { "abstract": "Catalysis of nucleon decay in white dwarfs is used to constrain the abundance of magnetic monopoles arising from Grand Unified Theories. Recent discoveries of the dimmest white dwarf ever observed, WD 1136-286 with $L = 10^{-4.94} L_{\\odot}$, place limits on the monopole flux. An abundance of monopoles greater than the new bound would heat this star to a luminosity higher than what is observed. The new bound is $(F/$cm $^{-2}$ s$^{-1}$ sr$^{-1}$) $(\\sigma \\upsilon/10^{-28} {\\rm cm}^2) < 1.3 \\times 10^{-20} (\\upsilon/ 10^{-3}c)^2$, where $\\upsilon $ is the monopole velocity. The limit is improved by including the monopoles captured by the main-sequence progenitor of the white dwarf: $(F/$cm $^{-2}$ s$^{-1}$ sr$^{-1}$ ) $(\\sigma \\upsilon /10^{-28} {\\rm cm}^2) < 3.5(26) \\times 10^{-21}$ for $10^{17}$ ($10^{16}$) GeV monopoles. We also note that the dependence on monopole mass of flux bounds due to catalysis in neutron stars with main sequence accretion has previously been calculated incorrectly (previously the bound has been stated as $F (\\sigma \\upsilon/10^{-28} {\\rm cm}^2) < 10^{-28} $ cm $^{-2}$ s$^{-1}$ sr$^{-1}$). We show that the correct bounds are somewhat weaker for monopole mass other than $10^{17}$ GeV. ", "introduction": "The question of whether or not magnetic monopoles exist has intrigued theorists and experimentalists for a long time \\cite{[1]}. In 1974, t'Hooft \\cite{[2]} and Polyakov \\cite{[3]} independently showed that magnetic monopoles always appear as stable topological entities in any Grand Unified theory (GUT) that breaks down to electromagnetism. Hence, if Grand Unified theories are shown to be correct, monopoles of mass in the range $10^{15}$ - $10^{19}$ GeV should exist. Rubakov \\cite{[4]} and Callan \\cite{[5]} calculated that these monopoles catalyze nucleon decay with a cross section characteristic of strong interactions, $\\sigma \\upsilon \\approx 10^{-28}$ cm$^2$. The abundance of these monopoles is an open question. The Kibble mechanism predicts roughly one monopole per horizon volume at the time of the Grand Unified phase transition. However, this estimate provides a severe overabundance of the number of monopoles: monopoles overclose the Universe by many orders of magnitude. Instead an inflationary epoch \\cite{[6]} may reduce their density in the Universe. Then the present abundance is difficult to estimate. A clue for experimentalists about what monopole flux to expect can be provided by astrophysics. The Parker bound \\cite{[7]} on the flux of monopoles was obtained by requiring survival of $\\mu$G magnetic fields observed in our Galaxy and gave $F \\leq 10^{-16}$ cm$^{-2}$ sr$^{-1}$ sec$^{-1}$. Subsequent improvements on this work include consideration of the monopole velocities \\cite{[8]} due to acceleration by the galactic magnetic field. Another improvement is the extended Parker bound, which required survival of a smaller seed magnetic field in the early period of the Galaxy \\cite{[9]}: $F \\leq 1.2 \\times 10^{-16} (\\frac {m}{10^{17}GeV})$ cm$^{-2}$ s$^{-1}$ sr$^{-1}$. Another class of methods for determination of the monopole flux is based on the hypothesis that GUT monopoles give rise to the catalysis of nucleon decay. The basic idea is that monopoles traveling through the Galaxy lose enough energy to be captured in an object (e.q. white dwarfs, neutron stars, etc.) where they subsequently catalyze nucleon decay. The energy produced by the nucleon decay heats up the object and results in a flux of photons from the surface of the object. One can then compare this predicted luminosity with what is actually observed. One must ensure that the monopoles would not make the object brighter than what is seen. The coolest star (or other object) seen provides the tightest limit on the monopole flux. If there were more monopoles than allowed by the bound, then the dimmest star observed could not exist. Several authors have carried out this kind of analysis in neutron stars \\cite{[10]}, nearby pulsar and white dwarfs. The strongest bound was obtained from consideration of the catalysis process in PSR 1929+10, an old pulsar \\cite{[11]}. From this pulsar, the bound on the product of monopole flux times cross section for catalysis is $(F/$cm$^{-2}$sr$^{-1}$sec$^{-1})(\\sigma\\upsilon / 10^{-28}$ cm$^{2}) \\leq 7 \\times 10^{-22}$. If one includes the monopoles captured by the main sequence progenitor of the white dwarf, this bound becomes even tighter \\cite{[12]}, \\hfill\\break $(F/$cm$^{-2}$ sr$^{-1}$ sec$^{-1})(\\sigma\\upsilon / 10^{-28}$ cm$^2$ $) \\leq 10^{-28}$. The consideration of monopole dynamics inside superconducting neutron-star cores leads to a bound $5 \\times 10^{-24} \\tau_{10}^{-2}$ cm$^{-2}$ sr$^{-1}$ s$^{-1}$ \\cite{[13]}, where $\\tau_{10}$ is the age (in $10^{10}$ years) of the pulsar's present magnetic field. As neutron stars are the densest astrophysical objects observed, they give rise to the tightest catalysis bounds. However, there is a certain amount of uncertainty due to the fact that the interiors of neutron stars are not well understood. For example, neutron stars can have very large magnetic fields $\\sim 10^{12}$G of unknown topology, and the motion of magnetic monopoles inside the neutron star would undoubtedly be affected by these magnetic fields. In addition neutron star interiors may contain pion condensates, again with uncertain effects on the monopoles. Because of the uncertainties with neutron star interiors, we turn to the next densest astrophysical objects in the Universe, white dwarfs. These stellar remnants are far better understood. The flux limits obtained from consideration of the catalysis process in white dwarfs are therefore important. Previously Freese \\cite{[14]} considered monopole catalyzed nucleon decay in white dwarfs. By comparing with the lowest luminosity white dwarf that had been seen at that time, she obtained a limit \\begin{equation} (F/{\\rm cm}^{-2}{\\rm s}^{-1}{\\rm sr}^{-1}) (\\sigma\\upsilon /10^{-28}{\\rm cm}^2) \\leq 2\\times 10^{-18}. \\end{equation} The present work is motivated by new observational data of cool white dwarfs \\cite{[15]}. In particular, Bergeron, Ruiz, and Leggett found a white dwarf 1136-286 (ESO 439-26) with luminosity $10^{-4.94}L_{\\odot}$; this is the dimmest white dwarf observed to date. We use the measured luminosities of old white dwarfs to constrain the radiation due to monopole-catalyzed nucleon decay and thus to obtain an upper limit to the average flux of monopoles in the Galaxy. Since a white dwarf with luminosity $10^{-4.94}L_{\\odot}$ is observed today, we know that the monopole-induced contribution to the white dwarf luminosity cannot exceed this value. These new data improve the limit on the monopole abundance due to catalysis in white dwarfs \\cite{[14]} by roughly two orders of magnitude. Of course, as dimmer white dwarfs are found, the bound will continue to get more restrictive. A monopole flux saturating this bound would keep the white dwarfs at luminosities at least this great and would lead to the prediction that no cooler white dwarfs will be found. As we will discuss, if it were indeed true that monopoles are keeping dwarfs hot, one would expect a different dependence of white dwarf luminosity on mass than expected in the standard model. We shall explicitly indicate the dependence of our results on various parameters. We will parametrize the properties of the white dwarf in terms of typical values from observations: for the mass, $M = M_{0.6} 0.6 M_{\\odot}$, for the radius $R = R_{9} 9\\times 10^8$cm, and for the average density $\\bar \\rho = 4\\times 10^5$ g cm$^{-3} M_{0.6} R_{9}^{-3}$. The central density is about an order of magnitude higher, $\\rho_{c} = 3 \\times 10^6$ g cm$^{-3} M_{0.6} R_{9}^{-3}$. Rubakov \\cite{[4]} estimated the product of cross section for catalysis and relative velocity $\\upsilon$ of the monopole and nucleon to be constant: $\\sigma \\upsilon = \\sigma_{0} = 10^{-28} ({\\sigma\\upsilon}_{-28})$ cm$^2$. (Throughout, we take $\\hbar = k_{B} = c = 1$.) For the thermal nucleon velocities expected inside a carbon and oxygen white dwarfs, $\\upsilon \\approx 10^{-3}c$, suppression effects may reduce the cross section by a factor of $10^{-2} s_{-2}$ \\cite{[16]}, and so we include this factor. In white dwarfs made of helium the suppression effects would be less effective $(s_{-2} = 10)$, and all the monopole flux bounds would be an order of magnitude stronger. ", "conclusions": "Figure 1 shows a plot of several monopole bounds: the Parker bound, the extended Parker bound, neutron star bounds, and the new white dwarf bound with and without main sequence capture. In the plots we have used the Wood cooling curves to be conservative. We have found that consideration of newly observed white dwarf 1136-286 with luminosity $10^{-4.94} L_\\odot$ and with new calculations of white dwarf cooling curves leads to a bound on the monopole flux that is two orders of magnitude lower than previous bounds due to catalysis in white dwarfs. The new bound is $F (\\sigma \\upsilon/10^{-28} {\\rm cm}^2) < 1.3 (1.9) \\times 10^{-20} (v/10^{-3}c)^2$ cm $^{-2}$ s$^{-1}$ sr$^{-1}$ for the Segretain \\cite{[19]} (Wood \\cite{[20]}) cooling curves respectively, where $\\upsilon $ is the monopole velocity. The limit is improved by including the monopoles captured by the main-sequence progenitor of the white dwarf: $F (\\sigma \\upsilon /10^{-28} {\\rm cm}^2) < 3.5(26) \\times 10^{-21}$ cm$^{-2}$ s$^{-1}$ sr$^{-1}$ for $10^{17}$ ($10^{16}$) GeV monopoles with $g=g_D$. Flux bounds for other monopole masses and parameters are given in Table 1. If cooler white dwarfs are discovered, a stricter bound on the monopole flux will result. We also showed that the dependence on monopole mass of flux bounds due to catalysis in neutron stars with main sequence accretion has previously been calculated incorrectly. Previously the bound due to catalysis in PSR 1929+10 with main sequence accretion has been stated as \\cite{[11]} $F (\\sigma \\upsilon/10^{-28} {\\rm cm}^2) < 10^{-28} $ cm $^{-2}$ s$^{-1}$ sr$^{-1}$. Instead, as can be seen in Table 1 and Figure 1, the correct bounds are somewhat weaker for monopole mass other than $10^{17}$ GeV. {\\bf Figure Caption} Bounds on the monopole flux as a function of monopole mass. The Parker bound \\cite{[7]} due to survival of the galactic magnetic field is plotted, as is the extended Parker bound \\cite{[9]} due to survival of the magnetic field early in the history of the Galaxy. Mass density limits ($\\Omega h^2 <1$) are plotted for a uniform density of monopoles in the universe. The bounds due to catalysis in white dwarf WD1136-286 as discussed in this paper are plotted; the plots assume the cooling curves of Wood \\cite{[20]}, and are very similar to those obtained using cooling curves of Segretain ${\\it et al.}$ In addition, the bounds from this white dwarf with main sequence accretion (WD/MS) are plotted for $g=g_D$ (solid line) and for $g=2 g_D$ (dotted line). The bounds due to calaysis in neutron star PSR 1929+10 are plotted, as are bounds due to this neutron star with main sequence accretion. Again the solid line is for $g=g_D$ and the dotted line is for $g=2g_D$. Note that the neutron star bounds with main sequence accretion have dependence on the monopole mass. \\vfill\\eject \\begin{center} \\begin{tabular}{|c|c|c|c|c|c|c|c|}\\hline &&&&&&$WD {\\rm w}/MS$&$NS {\\rm w}/MS$\\\\ \\cline{7-8} $M_m$ (GeV)&$\\beta$&${b_{\\rm crit}\\over R}$&$g/g_D$ &$N_{MS}/10^{38}F$&$N_{WD}/10^{38}F$&$F(\\sigma\\upsilon)_{-28}$&$ F(\\sigma\\upsilon)_{-28}$ \\\\[5pt] \\hline \\lower1.5ex\\hbox{$10^{15}$}&\\lower1.5ex\\hbox{$10^{-2}$} &{0.4}&{1}&2.5 &\\lower1.5ex\\hbox{0.17} &$8.2\\times 10^{-20}$& $6.2\\times 10^{-27}$ \\\\ \\cline{3-5}\\cline{7-8} &&0.56&{2}&4.9 && $4.3\\times 10^{-20}$& $3.2\\times 10^{-27}$ \\\\\\hline \\lower1.5ex\\hbox{$10^{16}$}&\\lower1.5ex\\hbox{$3\\times 10^{-3}$}&0.48&1&7.4&\\lower1.5ex\\hbox{1.8}&$2.4\\times 10^{-20}$& $2.1\\times 10^{-27}$ \\\\ \\cline{3-5}\\cline{7-8} &&0.62&2&12.3 && $1.6\\times 10^{-20}$& $1.3\\times 10^{-27}$ \\\\\\hline \\lower1.5ex\\hbox{$10^{17}$}&\\lower1.5ex\\hbox{$10^{-3}$}&0.54&1&52&\\lower1.5ex\\hbox{17}& $3.2\\times 10^{-21}$& $3.0\\times 10^{-28}$ \\\\ \\cline{3-5}\\cline{7-8} &&0.68&2&82 &&$2.2 \\times 10^{-21}$& $1.9\\times 10^{-28}$ \\\\ \\hline \\lower1.5ex\\hbox{$10^{18}$}&\\lower1.5ex\\hbox{$10^{-3}$}&---&1&---& \\lower1.5ex\\hbox{17 }&---&---\\\\ \\cline{3-5}\\cline{7-8}&&0.24&2&10 &&$8.1 \\times 10^{-21}$& $1.6\\times 10^{-27}$ \\\\\\hline \\end{tabular} \\end{center} \\bigskip \\begin{center} \\parbox{6in}{\\small Table 1: Bounds on the flux $F$ of magnetic monopoles in cm$^{-2}$s$^{-1}$sr$^{-1}$. Monopoles captured by white dwarfs (WD) or neutron stars (NS) and their main sequence (MS) progenitors are included. The white dwarf cooling time is taken to be $\\tau = 9.63$ Gyr. $M_m$ is the monopole mass in GeV, $\\beta$ is the monopole velocity, ${b_{\\rm crit} \\over R}$ is the ratio of the critical impact parameter for a monopole in units of the radius of the main sequence star, and the monopole charge is $g= 69 e (g/g_D)$ in units of the Dirac charge $g_D$. The number of monopoles captured by the MS progenitor and by the white dwarf are $N_{MS}$ and $N_{WD}$ respectively. The second to last column is the flux bound due to catalysis in WD 1136-286 (with MS monopoles included). The last column is the (corrected) flux bound due to catalysis in neutron star PSR 1929+10 (with MS monopoles included).} \\end{center} \\vspace{10mm} \\begin{center} \\begin{tabular}{|c|c|c|c|c|c|c|}\\hline $M_m$ (GeV)&$\\beta$&${b_{\\rm crit}\\over R}$&$g/g_D$&$N_{MS}/10^{38}F$&$N_{WD}/10^{38}F$&$F(\\sigma\\upsilon)_{-28}/ {\\rm cm}^{-2} {\\rm s}^{-1} {\\rm sr}^{-1}$ \\\\[5pt] \\hline \\lower1.5ex\\hbox{$10^{15}$}&\\lower1.5ex\\hbox{$10^{-2}$} &{0.4}&{1}&2.5 &\\lower1.5ex\\hbox{0.11} &$8.4\\times 10^{-20}$ \\\\ \\cline{3-5}\\cline{7-7} &&0.56&2&4.9 && $4.4\\times 10^{-20}$ \\\\\\hline \\lower1.5ex\\hbox{$10^{16}$}&\\lower1.5ex\\hbox{$3\\times 10^{-3}$}&0.48&1&7.4 &\\lower1.5ex\\hbox{1.2}&$2.6\\times 10^{-20}$ \\\\ \\cline{3-5}\\cline{7-7} &&0.62&2&12.3 && $1.6\\times 10^{-20}$ \\\\\\hline \\lower1.5ex\\hbox{$10^{17}$}&\\lower1.5ex\\hbox{$10^{-3}$}&0.54&1&52 &\\lower1.5ex\\hbox{11 }& $3.5\\times 10^{-21}$ \\\\ \\cline{3-5}\\cline{7-7} &&0.68&2&82 &&$2.4 \\times 10^{-21}$ \\\\ \\hline \\lower1.5ex\\hbox{$10^{18}$}&\\lower1.5ex\\hbox{$10^{-3}$}&---&1&---& \\lower1.5ex\\hbox{11 }&--- \\\\ \\cline{3-5}\\cline{7-7}&&0.24&2&10 &&$1.0 \\times 10^{-20}$ \\\\\\hline \\end{tabular} \\end{center} \\bigskip \\begin{center} {\\small Table 2: Same as table 1 for white dwarfs, but for cooling time $\\tau = 6.47$ Gyr.} \\end{center} {\\bf Acknowledgements:} We thank J. Allyn Smith, D. Graff, G. Laughlin, G. Tarle, and V.D. Ivanov for helpful conversations. We acknowledge support from the DOE at the University of Michigan." }, "9804/astro-ph9804238_arXiv.txt": { "abstract": "s{ We use numerical simulations of ray tracing through N-body simulations to investigate weak lensing by large-scale structure. These are needed for testing the analytic predictions of two-point correlators, to set error estimates on them and to investigate nonlinear gravitational effects in the weak lensing maps. On scales larger than 1 degree gaussian statistics suffice and can be used to estimate the sampling, noise and aliasing errors on the measured power spectrum. For this case we describe a minimum variance inversion procedure from the 2-d to 3-d power spectrum and discuss a sparse sampling strategy which optimizes the signal to noise on the power spectrum. On degree scales and smaller the shear and convergence statistics lie in the nonlinear regime and have a non-gaussian distribution. For this regime ray tracing simulations are useful to provide reliable error estimates and calibration of the measurements. We show how the skewness and kurtosis can in principle be used to probe the mean density in the universe, but are sensitive to sampling errors and require large observed areas. The probability distribution function is likely to be more useful as a tool to investigate nonlinear effects. In particular, it shows striking differences between models with different values of the mean density $\\Omega_m$. } ", "introduction": "Weak lensing by large-scale structure (LSS) shears the images of distant galaxies. The first calculations of weak lensing by LSS (Blandford et al. 1991; Miralda-Escude 1991; Kaiser 1992), based on the pioneering work of Gunn (1967), showed that lensing would induce coherent ellipticities of order 1$\\%$ over regions of order one degree on the sky. Recently several authors have extended this work to probe semi-analytically the possibility of measuring the mass power spectrum and cosmological parameters from the second and third moments of the induced ellipticity or convergence (e.g. Bernardeau et al. 1997; Jain and Seljak 1997; Schneider et al. 1997). The analytical work cited above suggested that nonlinear evolution of the density perturbations that provide the lensing effect can significantly alter the predicted signal. It enhances the power spectrum on scales below one degree and makes the probability distribution function (pdf) of the ellipticity and convergence non-Gaussian. We have carried out numerical simulations of ray tracing through N-body simulation data to compute the fully nonlinear moments and pdf. Details of the method and results are presented in a forthcoming paper; here we summarize the method and present some highlights of the results in Figures 1-5. We also discuss reconstruction of the dark matter power spectrum and error estimation using the gaussian approximation. ", "conclusions": "" }, "9804/astro-ph9804242_arXiv.txt": { "abstract": "The spatial distribution of the Ly$\\alpha$ forest is studied using new HST data for the quasar pair Q~1026--0045 A and B at $z_{\\rm em}$~=~1.438 and 1.520 respectively. The angular separation is 36~arcsec and corresponds to transverse linear separations between lines of sight of $\\sim$300$h^{-1}_{50}$~kpc ($q_{\\rm o}$~=~0.5) over the redshift range 0.833~$<$~$z$~$<$~1.438. From the observed numbers of coincident and anti-coincident Ly$\\alpha$ absorption lines, we conclude that, at this redshift, the Ly$\\alpha$ structures have typical dimensions of $\\sim$500$h^{-1}_{50}$~kpc, larger than the mean separation of the two lines of sight. The velocity difference, $\\Delta V$, between coincident lines is surprisingly small (4 and 8 pairs with $\\Delta V$~$<$~50 and 200~km~s$^{-1}$ respectively). \\\\ Metal line systems are present at $z_{\\rm abs}$~=~1.2651 and 1.2969 in A, $z_{\\rm abs}$~=~0.6320, 0.7090, 1.2651 and 1.4844 in B. In addition we tentatively identify a weak Mg~{\\sc ii} system at $z_{\\rm abs}$~=~0.11 in B. It is remarkable that the $z_{\\rm abs}$~=~1.2651 system is common to both lines of sight. The system at $z_{\\rm abs}$~=~1.4844 has strong O~{\\sc vi} absorption.\\\\ There is a metal-poor associated system at $z_{\\rm abs}$~=~1.4420 along the line of sight to A with complex velocity profile. We detect a strong Ly$\\alpha$ absorption along the line of sight to B redshifted by only 300~km~s$^{-1}$ relatively to the associated system. It is tempting to interpret this as the presence of a disk of radius larger than 300$h^{-1}_{50}$~kpc surrounding quasar A. ", "introduction": "\\label{intr} One way to probe the transverse extension of the gaseous structures giving rise to the Ly$\\alpha$ forest seen in the spectrum of quasars is to observe multiple lines of sight to quasars with small angular separations on the sky and search the spectra for absorptions coincident in redshift. This technique originated with a suggestion by Oort (1981) to test the possibility that the Ly$\\alpha$ forest clouds originate in large pancake structures. The first discoveries of common and associated absorption using pairs of distinct quasars (with separations $\\sim$~1~arcmin) were made by Shaver et al. (1982) and Shaver \\& Robertson (1983). These already indicated the possible existence of very large absorber sizes (hundreds of kpc), even for the Ly$\\alpha$ clouds. At about the same time Sargent et al. (1982) found no detectable tendency for Ly$\\alpha$ lines to correlate in QSO pairs separated by a few arcmin. Spectra of pairs of gravitational lens images revealed common absorptions on smaller scales (Weyman \\& Foltz 1983, Foltz et al. 1984). The idea that Ly$\\alpha$ clouds might have large sizes remained controversial untill the analysis by Smette et al. (1992), later confirmed by Dinshaw et al. (1994), Bechtold et al. (1994), Crotts et al. (1994), Bechtold \\& Yee (1994), Smette et al. (1995), D'Odorico et al. (1998). Recently, Dinshaw et al. (1995) derived a radius of 330$h^{-1}_{50}$ kpc at $z$~$\\sim$~0.7 for spherical clouds from observation of Q0107--0232 and Q0107--0235 separated by 86~arcsec. Larger separations have been investigated by Crotts \\& Fang (1997) and Williger et al. (1997). Both studies conclude that the clouds should be correlated on scales larger than 500~kpc.\\par\\noindent Here we present observations of Q1026--005~A ($m_{\\rm r}$~=~18.4, $z_{\\rm em}$~=~1.438) and B ($m_{\\rm r}$~=~18.5, $z_{\\rm em}$~=~1.520), two distinct quasars separated on the sky by 36~arcsec or 300~$h^{-1}_{50}$~kpc ($q_{\\rm o}$~=~0.5) at $z$~$\\sim$~1. ", "conclusions": "" }, "9804/astro-ph9804304_arXiv.txt": { "abstract": "How has the ``fluffy'' core of the Sgr dwarf galaxy survived multiple strong shocks from the tidal force of the Galactic halo and disc since the formation of the core a Hubble time ago? A scenario that Sgr was deflected to its current orbit by the Magellanic Clouds after a rendezvous on the north Galactic pole $2-3$ Gyrs ago is examined. It is shown that the conditions of the collision fix both the sense of circulation of Sgr and the LMC around the Galaxy and the slope of the Galactic rotation curve. The model argues that the two orthogonal polar circles traced by a dozen or so Galactic halo dwarf galaxies and globular clusters (LMC-SMC-Magellanic Stream-Draco-Ursa Minor along $l \\approx 270^o$ and M54-Ter 7-Ter 8-Arp 2-NGC 2419-Pal 15 along $l \\approx 0^o$) are streams of tidal relics from two ancient galaxies which was captured on two intersecting polar rosette orbits by the Galaxy. Our results favor the interpretation of microlensing towards the LMC being due to source or lens stars in tidal features of the Magellanic Clouds. We discuss direct and indirect observations to test the collision scenario. ", "introduction": "The recently discovered dwarf galaxy at about $25$ kpc from the Sun in the direction of the Sagittarius constellation (Ibata, Gilmore \\& Irwin 1994) is the closest galaxy known to us. It is traced by two long trailing/leading tails on the sky (together more than $8^o\\times 22^o$ in solid angle) with most of its stars still clustered around a low density luminous core (roughly $0.001 L_\\odot{\\rm pc}^{-3}$ with semi-axes $1:1:3$ kpc). It is puzzling why this fluffy core of the dwarf galaxy has not been fully ``digested'' by the Galaxy, in the sense that stars have not fully dispersed out of the core despite the severe shocks at pericentric passage from the tidal force of the Galactic halo (about $10-100$ times stronger than that experienced by satellites in the outer halo, the Magellanic Clouds and the Fornax dwarf galaxy included) and shocks when crossing the disc of the Galaxy. The best fit to Sgr's morphology, radial velocity (Ibata, Gilmore \\& Irwin 1995) and proper motion (Ibata, Wyse, Gilmore \\& Suntzeff 1997) yields an orbit with a pericenter-to-pericenter period of about $0.8$ Gyr and a peri and apo-center at about 10 and 50 kpc respectively (Vel\\'azquez \\& White 1995). Simulations show that if a typical Galactic dwarf galaxy (such as Fornax) were replaced on Sgr's orbit, it would dissolve in no more than two peri-centric passages by the strong peri-centric tidal shock of the Galaxy near 10 kpc (Vel\\'azquez \\& White 1995; Johnston, Spergel \\& Hernquist 1995; Johnston, Hernquist \\& Bolte 1996; Edelsohn \\& Elmegreen 1997). This apparently contradicts the observation that the dominant stellar population in the core is older than 10 Gyrs (Mateo et al. 1995, Fahlman et al. 1996), implying that Sgr has survived $10-20$ peri-centric tidal shocks of the Galaxy. To circumvent this dilemma we need to abandon either or both of the following hidden assumptions: (i) the light distribution of Sgr traces its mass, and (ii) Sgr has always been on the same low-pericentric orbit in a rigid Galactic potential for the past 5 to 10 Gyrs. Ibata, Wyse, Gilmore \\& Suntzeff (1997) postulate a dense dark halo of Sgr surrounding the luminous part to hold the system together; they require Sgr's mass density to be uniform inside about 3 kpc of its core with a value ($\\sim 0.03 M_\\odot{\\rm pc}^{-3}$) several times the mean Galactic halo density inside $10$ kpc ($0.013M_\\odot{\\rm pc}^{-3}$). An inspection of Sgr's rosette-like orbit in relation to that of the Magellanic Clouds (MCs) offers a completely different line of thought. They are on nearly orthogonal planes intersecting along the poles with their Galactocentric radii overlapping at about 50 kpc. So an encounter at the north or south pole some time in the past or future is quite inevitable. A recent preprint by Ibata \\& Lewis (1998), shortly after the completion of the work reported in this {\\it Letter}, also remarked on a small chance of an interaction after noticing in their simulations a weak perturbation to Sgr's orbit when they turned on the moving gravitational field of the massive MCs. Unfortunately the effect was in the end neglected on grounds of low probability without thoroughly exploring the parameter space (of satellite velocities and the Galactic potential) and the important consequences of a rare strong interaction. So same as Ibata et al. (1997) they were thus left with no alternative but to conclude a massive dark halo of Sgr to be the only explanation for Sgr's presence on a low-pericentric orbit after a Hubble time. In this {\\it Letter} we examine the encounter scenario, as illustrated in Fig. 1, where Sgr has been pulled back from an originally high angular momentum/energy orbit to the present low angular momentum/energy orbit by the massive MCs. A recent encounter would have the advantage to allow Sgr to spend most of its lifetime on a ``safe'' orbit with a pericenter (say, $20$ kpc) too high to be harassed by the sharply declining tidal force of the Galaxy (e.g., Kroupa 1997, Oh, Lin \\& Aarseth 1995); in a halo with an $r^{-2}$ density profile the pericentric shock would drop by a factor of $4$ from $10$ kpc to $20$ kpc. \\onecolumn \\begin{figure} \\epsfysize=15cm \\centerline{\\epsfbox{sgrmc3d.ps}} \\caption{ A 3D view and $x-y$, $y-z$ projections of the orbits of the LMC/SMC (thick red/dashed green lines near the $x=0$ plane) and Sgr (thin blue curve near the $y=0$ plane); the Sun is be at $(x,y,z)=(-8,0,0)$ kpc. The three systems are integrated backward for $3$ Gyrs (with ellipses marking steps of $0.5$ Gyr) from the present epoch (with velocity vectors marked by arrows) inside the Galactic potential.\\label{sgrmc3d} } \\end{figure} \\twocolumn Various interesting aspects of this scenario will be discussed at the end. But the aim of this {\\it Letter} is to report an independent constraint on the rotation curve of the Galaxy as imposed purely by {\\it timing the collision}. The essence is the following. The random chance for the LMC and Sgr to meet each other is obviously low, about 1\\% for a 10 kpc closest approach in the past 3 Gyrs for a general set of Galactic potentials and initial conditions of the satellites. So the same argument could be inverted: once we accept the deflection by the MCs as a plausible way out of Sgr's dilemma a stringent set of conditions on the potential of the halo and the proper motions of the satellites must follow. ", "conclusions": "" }, "9804/astro-ph9804214_arXiv.txt": { "abstract": "A new approach to the study of the large-scale stellar cluster distribution in the Galaxy based on two-point correlation techniques is presented. The basic formalism for this method is outlined and its applications are then investigated by the use of a simple model of cluster distribution in the Galaxy. This provides an estimate of the potentials of the two-point correlation function for indicating clustering in the measured star positions, which can be related to the presence of star clusters in the observed field. This technique is then applied to several areas of the Two Micron Galactic Survey catalogue, from which information is obtained on the distribution of clusters according to position in the Galaxy, as well as about age, density of stars, etc. ", "introduction": "Open star clusters provide valuable information on the evolution of the Galaxy. In this paper two-point correlation techniques are used to analyse the distribution of open clusters in order to gain an insight into the structure and evolution of the Galaxy. Open-cluster distributions have been widely studied at optical wavelengths as a means of studying Galactic structure and evolution (see, for example, Lyng\\aa \\ 1987b; Janes \\& Phelps 1994). The Lyng\\aa\\ catalogue of open clusters (Lyng\\aa \\ 1987a) lists about 1200 clusters, which represent nearly all the open clusters accessible in the visible. Knowledge of the positions and ages of these clusters (a method of age determination for clusters is given by Carraro \\& Chiosi 1994) enables the scale length and scale height of the disc to be derived for both young and old clusters (for a review of old open clusters, see Friel 1995) and theories to be developed on their formation and destruction throughout the history of the Galaxy. The limitations on these studies are imposed by the maximum distance at which open clusters can be detected. Most of the cataloged open clusters are in the solar neighbourhood, and very few have distances greater than 3 kpc (Payne-Gaposchkin 1979). Hence, information is obtained only for a small region of the Milky Way. The problem is caused by interstellar extinction. An excellent tool for studying star clusters and star formation regions is to observe in the infrared (Wynn-Williams 1977), where the effect of the extinction is far less. To date, however, the infrared has been little used in this field owing the absence of suitable databases. The $K$ band is probably the best region of the spectrum for tracing the stellar distribution of the Galaxy. The $K$ radiation is a mass tracer in spiral galaxies because it follows the old stellar population (Rhoads 1995). Furthermore, the $K$ light is dominated by high-mass stars in star formation regions, i.e. in open clusters, so it is especially appropriate in the search for young clusters, which are rich in massive stars. As explained in more detail by Garz\\'on et al. (1993), the Two micron Galactic survey (TMGS) is a $K$-band survey of various regions along the Galactic equator between $-5^o0$ that the probability is greater than random, i.e. that there is clustering; and $\\xi (r)<0$ that the probability is less than random, i.e. that there is relative avoidance). In the same way, the two-point correlation function can be defined for two dimensions on the surface on to which the distribution is projected (the celestial sphere in the case considered here). This is called the two-point angular correlation function (TPACF) and is defined as \\begin{equation}\\omega (|\\vec{\\theta}-\\vec{\\theta}'|)=\\frac{\\langle\\sigma (\\vec{\\theta}) \\sigma(\\vec{\\theta}')\\rangle} {\\langle\\sigma(\\vec{\\theta})\\rangle^2} -1 \\label{omega},\\end{equation} where $\\sigma $ is the surface density per unit solid angle. Another mathematical technique for deciding whether a distribution is non-Poissonian is area tessellation, as was used by Bal\\'azs (1995) to test the grouping tendency of H$_\\alpha$-emission stars in the Orion molecular clouds without giving a quantitative measure of the departure from the Poissonian distribution. See also P\\'asztor et al. (1993), P\\'asztor \\& T\\'oth (1995) and references therein for other astronomical applications of spatial statistics. \\subsection{Relationship between the TPCF and the TPACF for stars} When applying the above definitions to stars in the Galaxy, the luminosity function and space density have to be taken into account. By generalizing the result of the Limber (1953) equation for constant density, the relationship between the TPCF (which is non-zero for distances less than $\\Delta r$) and the TPACF (for small $\\theta$) for any density distribution is \\[ \\omega _{\\rm t}(\\theta)\\approx \\frac{1}{\\langle\\sigma _{\\rm t}\\rangle^2} \\int _0^{\\infty }dr\\ r^4 \\langle\\rho \\rangle^2(r) \\int_{r-\\Delta r}^{r+\\Delta r}dr' \\]\\[ \\times \\int _{M_{\\rm min}(r)}^{M_{\\rm max}(r)} dM\\ \\phi(M)\\int_{M_{\\rm min}(r)}^{M_{\\rm max}(r)}dM'\\ \\phi(M') \\]\\begin{equation} \\times \\xi \\left(\\sqrt{ (r\\theta )^2+(r-r')^2};r,M,M'\\right) ,\\label{omegat2}\\end{equation} where $r$ is the distance along the line of sight, $M$ the absolute magnitude, $\\phi (M)$ the lu\\-mi\\-no\\-si\\-ty function and \\begin{equation} \\langle\\sigma _{\\rm t}\\rangle=\\int _0^{\\infty }dr\\ r^2\\langle\\rho \\rangle(r)\\int _{M_{\\rm min}(r)} ^{M_{\\rm max}(r)} dM\\ \\phi(M) .\\label{sigmat2}\\end{equation} The minimum and maximum values of $M$ for a distance $r$ depend on the minimum and maximum values of the apparent magnitude and the extinction along the line of sight. In this case, it is assumed that the absorption is not patchy, i.e. that it is independent of $\\theta $ for small angles. This is not exactly true but it will be show in Section \\ref{.extincorr} that the effects are negligible. The subscript `t' stands for `total', a projection of all distances and magnitudes, and $\\langle\\sigma _{\\rm t}\\rangle$ is the total two-dimensional density for all distances and magnitudes. In the literature, $\\langle\\sigma _{\\rm t}\\rangle$ is also called $A(m_{\\rm min},m_{\\rm max},l,b)$ and represents the star counts in the magnitude range (Bahcall 1986). This expression enables the TPACF to be found once the three-dimensional distribution of the stars is known and forms the basis of this article, in which we create a model distribution of the stars and compare the results obtained with those observationally in order to investigate the distribution of clustering in the structure of our Galaxy. In general, the TPACF cannot be inverted to give the TPCF due to the multiplicity of possible solutions and to the lack of precise knowledge of certain parameters. However, there are certain cases in which the equation can be inverted and TPCF obtained from TPACF (Fall \\& Tremaine 1977). A trivial example where inversion is possible is that of a Poissonian three-dimensional distribution, which implies a Poissonian projected distribution and vice versa, i.e. $\\xi =0$, $\\omega =0$ on all scales. Another example is when $\\langle\\rho\\rangle(r)$ is a constant independent of $r$. \\subsection{ Definition of new variables} In order to simplify the comparison of the level of clustering for different regions of the sky, two new variables will be introduced. $\\theta _{\\rm max}$ is defined as the first zero of $\\omega _{\\rm t}(\\theta)$. In this article (see for example Fig. \\ref{Fig:a052}), $\\omega _{\\rm t}$ is positive up to a separation $\\theta _{\\rm max}$. For values greater than $\\theta _{\\rm max}$ this value is small and oscillates about zero, as there is no correlation among stars separated by large angular distances. Another definition, corresponding to the integration of $\\omega _{\\rm t}$ up to the limit $\\theta =\\theta _{\\rm max}$ ($\\theta >\\theta _{\\rm max}$ would give a null contribution to the integral), is \\begin{equation} C_2\\equiv \\frac{\\int _0 ^{\\theta _{\\rm max}}d\\theta \\ \\theta \\omega _{\\rm t}(\\theta ) }{\\theta _{\\rm max}^2} ,\\label{C2}\\end{equation} which means the excess (when $C _2$ is positive) or deficit (when $C _2$ is negative) of the relative number of objects with respect to a Poissonian distribution in a circle centred on an arbitrary star on the celestial sphere, within the observed solid angle and with angular radius $\\theta _{\\rm max}$. The relative correlation within the angular scale $\\theta _{\\rm max}$ is therefore measured. We call $C_2$ the `accumulation parameter' (N.B. there are also other definitions in the literature of the TPACF integral, e.g. Wiedemann \\& Atmanspacher 1990). The variable $C_2$ has a clear meaning associated with projected clustering and is also a useful number to measure. Since it sums several values of $\\omega $ for different angles, it condenses the information of interest into a single number that can be compared for different samples of stars and give the degree of clustering. This parameter is a mathematical expression of the degree of clustering seen in fields of stars. The idea that we wish to stress here is that all mathematical developments described in this paper are designed to put the intuitive idea of clustering to a reliable test. These calculations are necessary for a quantitative, as opposed to a merely qualitative, description of clustering. Applying the expression (\\ref{omegat2}) of $\\omega _{\\rm t}$ in $C_2$, we get \\[ C_2=\\frac{1}{\\langle\\sigma _{\\rm t}\\rangle^2\\theta _{\\rm max}^2} \\int _0^{\\infty }dr\\ r^2\\langle\\rho (r)\\rangle^2 \\int _{M_{\\rm min}(r)}^{M_{\\rm max}(r)} dM\\ \\phi(M) \\]\\begin{equation} \\times \\int _{M_{\\rm min}(r)}^{M_{\\rm max}(r)} dM'\\ \\phi(M') \\int _0^{\\theta _{\\rm max}r} dy\\ y \\Xi (y;r,M,M') \\label{C2a},\\end{equation} where $\\Xi $, an integrated TPCF, is \\[ \\Xi(y;r,M,M') \\]\\begin{equation}= \\int_{r-\\Delta r}^{r+\\Delta r}dr' \\xi\\left (\\sqrt{y^2+(r-r')^2};r,M,M'\\right) .\\label{Xi}\\end{equation} \\subsection{ Further approximations} In order to simplify the above calculations, it will be assumed that the distribution of stars does not depend on their luminosity, i.e. that $\\xi (y;r,M,M')=\\xi (y;r)$. This is not completely true as there is a small dependence on the distribution of stars in a cluster according their masses, and the luminosities are dependent on the masses. A complete calculation taking the luminosity function into consideration would be of great value. However, the relationship between the TPCF and the luminosity function is uncertain, although the effects of this approximation for the detection of clusters are expected to be small. With this approximation, and from (\\ref{sigmat2}) and (\\ref{omegat2}), \\begin{equation} \\langle\\sigma _{\\rm t}\\rangle=\\int _0^{\\infty}dr\\ \\langle N^*\\rangle(r) \\end{equation} and \\begin{equation} \\omega _{\\rm t}(\\theta)=\\frac{1}{\\langle\\sigma _{\\rm t}\\rangle^2} \\int _0^{\\infty}dr\\ \\langle N^*\\rangle^2(r) \\Xi (r\\theta ;r) ,\\label{omegat*}\\end{equation} where $\\langle N^*(r)\\rangle$ is the number of stars observed per unit solid angle at a distance $r$: \\begin{equation} \\langle N ^*\\rangle(r)=r^2\\langle\\rho \\rangle(r) \\int _{M_{\\rm min}(r)}^{M_{\\rm max}(r)} dM\\ \\phi(M) .\\label{N*}\\end{equation} The variable $\\omega _{\\rm t}$ can also be expressed as \\begin{equation} \\omega _{\\rm t}(\\theta)=\\overline{\\Xi (\\overline{r}\\theta )} \\label{omegaavXi}\\end{equation} where the averages $\\overline{r}$ and $\\overline{\\Xi }$ are such that match (\\ref{omegat*}). Also, from (\\ref{C2a}), \\begin{equation} C_2=\\frac{1}{\\langle\\sigma _{\\rm t}\\rangle^2\\theta _{\\rm max}^2} \\int _0^{\\infty }dr\\ \\frac{\\langle N^* (r)\\rangle^2} {r^2} \\int _0^{\\theta _{\\rm max}r} dy\\ y \\Xi (y;r) .\\label{C2*}\\end{equation} \\noindent This last equivalence is a way of averaging the function $\\xi $. Hence, high values of $C_2$ indicate that there must be high projected clustering in the direction of the beam. \\subsection{Patchiness of extinction} \\label{.extincorr} It is clear that extinction can distort the observed counts, the amount of the distortion being a matter of controversy. It is generally accepted that in the optical wavelengths this influence is very severe, particularly in regions near to or in the Galactic plane in the inner Galaxy, where the strong and patchily distributed obscuration makes it difficult to penetrate deep into the Galaxy. The amount of extinction decreases substantially with increasing wavelength. Maihara et al. (1978) quoted a value of 0.17 mag kpc$^{-1}$ as typical for extinction in the Galactic plane in the $K$ band, compared with 1.9 mag kpc$^{-1}$ for the $V$ band (Allen 1973). This has two important consequences. First, the $K$ band is more effective at penetrating the interstellar dust. Secondly, the observed stellar dis\\-tri\\-bu\\-tion more closely resembles the true distribution. For the second argument to be true it is necessary that the obscuration in the $K$ band should not only be smaller in amount than in the $V$ band, but also that its patchiness should be less important. This rather uniform distribution of the interstellar extinction in $K$ can be inferred from the TMGS histograms in several cuts across the Galactic plane. Garz\\'on et al. (1993, their Fig. 8) compared the observed stellar distribution in the TMGS and the GSC in the $V$ band. It is noticeable how uniform the $K$ histograms are, particularly when compared with those for the GSC. Except for small portions highly concentrated in the Galactic plane and more marked in the central regions, the shape of the high spatial resolution distribution curves of the TMGS does not exhibit the `noisy' pattern of the GSC plots, which is certainly due to the presence of strong and patchily distributed extinction. Hammersley et al. (1994) showed similar histograms for different areas which also have similar shapes. Moreover, a good fit to a classical ex\\-po\\-nen\\-tial disc can be seen in Fig. 3 of that paper; this would not be the case if the extinction were important and non-uniform. This conclusion can also be reached from the contour maps of the bulge of the Galaxy of Dwek et al (1995), who showed the residuals of the DIRBE data after disc subtraction and extinction correction. Again, the general shape of the maps proves the basic uniformity of extinction distribution in the near infrared. We now estimate these effects. From (\\ref{sigmat2}) with the change of variable $r=10^{(5+m_{\\rm max}-M_{\\rm max})/5}$ and \\begin{equation} \\Phi (M_{\\rm max})=\\int _{-\\infty} ^{M_{\\rm max}} dM\\ \\phi(M) \\approx \\int _{M_{\\rm min}}^{M_{\\rm max}} dM\\ \\phi(M) ,\\end{equation} the local cumulative counts $\\sigma _{\\rm t}$ follow the expression \\[ \\sigma _{\\rm t}=\\langle\\sigma _{\\rm t}\\rangle _{local} =200(\\ln\\ 10)10^{3m_{\\rm max}/5}\\omega \\int_{-\\infty }^\\infty dM_{\\rm max} \\]\\begin{equation}\\times D\\left( 10^{(5+m_{\\rm max}-M_{\\rm max})/5}\\right) 10^{-3M_{\\rm max}/5} \\Phi (M_{\\rm max}), \\end{equation} ignoring the variation of extinction with the distance. If we take the density $D$ as constant, then \\begin{equation} \\sigma _{\\rm t}= N(m_{\\rm max})\\propto 10^{3m_{\\rm max}/5} .\\end{equation} Taking $D$ as constant is sufficient for estimating the the order of magnitude of the patchiness due to extinction. In any case, the above proportionality is followed in the observed cumulative counts but with a constant value of between 1 and 2 instead of $3/5$ in the exponent. An excess of extinction, $\\Delta a(\\theta)$, due for example to a cloud at an angular distance $|\\vec{\\theta } -\\vec{\\theta _0}|$ with respect to a given point $\\theta _0$, will cause a reduction in the apparent flux of a fraction, $f$, of stars (behind the cloud), thereby creating the same effect as a reduction in the maximum apparent magnitudes of these stars by $\\Delta a(\\theta)$ mag, or, if $\\Delta a(\\theta)$ is relatively small, a reduction in $m_{\\rm max}$ by $f\\Delta a(\\theta)$ mag for all stars. Hence, \\begin{equation} \\sigma _{\\rm t}(\\theta)\\sim \\sigma _{\\rm t} (\\theta _0)10^{-3(a(\\theta )-a(\\theta _0))f/5} \\label{sigmada} .\\end{equation} If it is assumed that the observed flux fluctuations, $\\Delta F$, are due mainly to extinction variations, with the small-fluctuation approximation, then both are related by \\begin{equation} \\Delta a=-2.5\\log \\left( 1-\\frac{1}{f}\\frac{\\Delta F}{F} \\right)\\approx \\frac{5\\log _{10}e}{2f}\\frac{\\Delta F}{F} \\end{equation} (the factor $f$ appears again here for the same reasons as above). So, from equation (\\ref{sigmada}), using the small-fluctuation approximation, \\begin{equation} \\frac{\\sigma _{\\rm t}(\\theta )} {\\sigma _{\\rm t} (\\theta _0)} \\sim \\frac{3}{2} \\frac{F(\\theta)}{F(\\theta _0)} \\label{omegaext} .\\end{equation} This means that the angular correlation of star density is about 3/2 times the angular correlation of the flux. Averaging the DIRBE $K$ flux (Boggess et al. 1992) fluctuations from the maps with $2520''$ resolution over $\\mid b\\mid \\le 3^\\circ$ for constant-$l$ strips over the range $-35^\\circ < l< 35^\\circ $ (where the effects of extinction are most relevant), we get root mean squares of \\begin{equation} 0.03< \\sigma (\\Delta F/F)_{l={\\rm const.}, \\mid b \\mid \\le 3^\\circ} < 0.23 ,\\end{equation} with an average of \\begin{equation} \\overline{\\sigma (\\Delta F/F) _{l={\\rm const.}, \\mid b \\mid \\le 3^\\circ}}= 0.10 .\\end{equation} The oscillations of flux fluctuations are not very high in the plane, their maximum being $2.3$ times the average. From equation (\\ref{omegaext}), and taking into account that the root mean square is $\\sqrt{\\omega (0)}$, \\begin{equation} \\overline{\\omega (0) _{l={\\rm const.}, \\mid b \\mid \\le 3^\\circ}} \\sim 0.015 \\end{equation} for regions of 2520$''$ in size. In the most unfavourable case, where the extinction is highest (multiplied by a factor of 2.3$^2$ because the maximum root mean square is $2.3$ times greater than the average), $\\omega (0) \\sim 0.08$. Higher-resolution flux maps are not available in the $K$-band for the whole sky so we cannot derive these numbers for smaller scales, but they are not expected to be much higher since average cloud size is of the order of degrees (rather higher than 42$'$) and the cloud distribution is fairly smooth. A fractal distribution would increase the contamination but this may apply only to very cold gas clouds (Pfenninger \\& Combes 1994) which are not the main cause of extinction in the $K$ band. We conclude that extinction in $K$ cannot be responsible for correlations $\\omega (0)$ greater than $\\sim 0.08$. This is just an estimate, but the order of magnitude should not be very different. As will be shown, the results when applied to the TMGS are above this value (see, for example, Fig. \\ref{Fig:a052}), and causes other than patchy extinction must explain this. ", "conclusions": "A technique is developed for searching for clustering in stellar surveys using correlation functions. The mathematical tools are useful for any field of stars and can be applied to any survey, especially those at carried out at infrared wavelengths, which permit a study of the distribution of stars throughout almost the entire Galaxy. The DENIS (Epchtein 1997) or 2MASS (Skrutskie et al. 1997) surveys will be ideal for this technique as the increased numbers of stars will reduce the errors. It is even possible, with a large number of stars in the survey, to apply the technique for different ranges of apparent magnitude. Studying the clustering of stars at different apparent magnitudes is equivalent to do studying in three dimensions ($l$, $b$ and the average distance $\\overline{r}$ which is associated with the treated range of magnitudes). A simple model has been developed. This model could be improved by introducing a density dependence as a function of the distance from the centre of the cluster, perhaps a power-law dependence. In this paper the method has been applied to the TMGS. Is has been shown that a simple model in which old open clusters trace the whole Galaxy with a density of clusters proportional to the density of stars agrees quite well with the data. An exception to the general agreement are specific regions in the plane where the higher-than-expected clustering can be a\\-ttri\\-bu\\-ted to star formation in the spiral arms. A second departure from the simple model is the reduced $C_2$ in the outer disc and in the bulge due to a lack of young clusters. In one of the regions with an excess, at $l=70^\\circ $ in the plane, the approximate limits for the cluster density and the density of stars inside the cluster are derived. These are, respectively, $5\\times 10^{-7}\\ {\\rm pc}^{-3} <{\\rm \\langle}n_{\\rm cl}{\\rm \\rangle}< 2 \\times 10^{-6}\\ {\\rm pc}^{-3}$ and $140\\ {\\rm pc}^{-3} < \\rho _{\\rm cl} < 700\\ {\\rm pc}^{-3}$. There is, however, a lower-than-expected correlation at $l=27^\\circ $, $b=0^\\circ $. There is believed to be a huge star formation region in this direction and the lack of correlation could be due to the star formation region being far larger than the sample area. As has been pointed out by Friel (1995), the oldest open clusters may be viewed from two perspectives with regard the formation of the Galaxy: a halo collapse or a continuous accretion and infall of material from the halo on to the Galactic disc. Either perspective is possible. The first should justify which were the original star formation regions that were the origin of the present old clusters in the outer disc and how they travelled there from their place of origin. The second perspective needs to test the infall of matter from the halo as well as the existence of star systems in the halo. Further improvements on these cluster searches and better numbers will give us a hint concerning these questions on the origin of the Galaxy. A better determination of $C_2$ in the bulge region will tell us about the age of bulge clusters if these exist. In this article we have observed a relative absence of correlation in the bulge that is somewhat less than the prediction of our simple model, but at best the prediction could say, as in the case of the anticentre, whether the correlation is greater or less than the improved model and enable us to reach further conclusions. \\subsection*" }, "9804/astro-ph9804164_arXiv.txt": { "abstract": "We analyze ultraviolet ($\\sim 1500$~\\AA) images of the old open clusters M67, NGC 188, and NGC 6791 obtained with Ultraviolet Imaging Telescope (UIT) during the second flight of the {\\em Astro} observatory in March 1995. Twenty stars are detected on the UIT image of M67, including 11 blue stragglers, seven white dwarf candidates, and the yellow giant -- white dwarf binary S1040. The ultraviolet photometry of the blue stragglers F90 (S975) and F131 (S1082) suggests that these stars have hot subluminous companions. We present a semi-empirical integrated ultraviolet spectrum of M67, and show that the blue stragglers dominate the integrated spectrum at wavelengths shorter than 2600~\\AA. The number of white dwarfs in M67 is roughly consistent with the number expected from white dwarf cooling models. Eight candidate sdB/sdO stars are detected in NGC 6791, and two are detected in NGC 188. The luminosity range $1.10 < \\log$ L/\\lsun\\ $< 1.27$, derived from the ultraviolet photometry of the six sdB candidates, is consistent with theoretical models of metal-rich hot horizontal branch (HB) stars. The fraction of hot HB stars in both NGC 6791 and NGC 188 is about 30\\%, implying that the integrated spectra of both clusters should show a UV turnup at least as strong as that observed in any elliptical galaxy. ", "introduction": "As a stellar population ages, the main-sequence turnoff becomes cooler and fainter, until, after an age of about 2 Gyrs, the ultraviolet ($\\sim$ 1500~\\AA) flux is no longer dominated by stars near the main-sequence, but instead must be due to a more exotic population of hot stars, such as blue stragglers, hot subdwarf (sdB, sdO) stars, hot post-asymptotic giant branch (post-AGB) stars, white dwarfs, or interacting binaries. The blue stragglers and hot subdwarfs are of particular interest since they might have significant roles, respectively, in two current problems in extragalactic astronomy: the age-dating of old galaxies from their rest-frame near-UV ($\\sim 2600$~\\AA) spectra, and understanding the origin of the observed UV-upturn in elliptical galaxies. In the absence of blue stragglers, the rest-frame near-UV spectrum of an old galaxy should be dominated by the turnoff population, and thus fitting the spectrum to population synthesis models should be a reliable method for age-dating the galaxy, although metallicity effects also need to be properly disentangled (\\cite{heap98}, \\cite{bruz97}). However, as noted by \\markcite{spin97}Spinrad et al. (1997), a significant underestimate of the galaxy age could result if blue stragglers contribute a large fraction of the galaxy near-UV flux, but are not included in the adopted population synthesis model. \\markcite{greg90}Greggio \\& Renzini (1990) outlined the various candidates for the origin of the observed UV-upturn in elliptical galaxies, which is characterized by a rising UV flux shortward of 1800~\\AA\\ (\\cite{bica96b}, \\cite{brown97}). Among these candidates, white dwarfs and hot post-AGB stars have the advantage of being relatively well-understood phases of stellar evolution, but lack sufficient fuel to explain the strongest UV-upturn galaxies. The more promising candidates are currently the extreme horizontal-branch (EHB) stars and their hot progeny; the AGB-manqu\\'e stars, which miss the AGB phase entirely, and the post-early AGB stars, which leave the AGB before the thermally pulsing phase. Although the horizontal branch (HB) morphology of the globular clusters becomes generally redder with increasing metallicity, a significant EHB population is theoretically expected at metallicities above solar (and an age larger than about 5 Gyr), if reasonable assumptions are made concerning the metallicity dependence of the mass loss rate and helium abundance (\\cite{dorm95}, \\cite{yi97a}). Spectroscopically, EHB and AGB-manqu\\'e stars are expected to appear, respectively, as sdB and sdO stars, and the measured effective temperatures and gravities of the field hot subdwarf stars is consistent with this scenario (\\cite{saff94}). However, the field hot subdwarfs apparently have a large binary fraction (\\cite{theis95}, \\cite{green98b}), which instead suggests that binary interactions might be involved in their formation (\\cite{bailyn95}). Despite their probable contributions to the ultraviolet spectra of galaxies, blue stragglers and hot subdwarfs are either neglected or poorly constrained in existing population synthesis models of old stellar populations. The currently favored explanations for the origin of blue stragglers involve binary interactions, either through mass-transfer processes, binary mergers, or direct stellar collisions (\\cite{bailyn95}, \\cite{leonard96}), but the large parameter space of these binary processes limits their inclusion in synthesis models to special cases (e.g.\\ \\cite{pols94}). Some attempts have been made to include EHB stars in synthesis models of metal-rich populations, but these are hampered by large uncertainties, for example, in the dependence of mass loss and helium abundance on metallicity (\\cite{bress94}, \\cite{yi97b}). The modeling of galaxy ultraviolet spectra should clearly benefit from empirical studies of the ultraviolet content of resolved, old, metal-rich stellar populations. Here we report on ultraviolet ($\\sim$ 1600~\\AA) images of the three old ($> 4$ Gyr) open clusters M67, NGC 188, and NGC 6791, which were obtained in March 1995 using the {\\em Ultraviolet Imaging Telescope} (UIT, \\cite{stech97}) during the second flight of the {\\em Astro} observatory. The $40'$ diameter field of view of UIT is sufficiently large to allow a complete census of the hot stellar population in the observed clusters. Thus, the UIT images of these three clusters can be used for an empirical study of the ultraviolet content of old, metal-rich stellar populations. As it turns out, M67 contains blue stragglers, but no hot subdwarfs, whereas NGC 188 and NGC 6791 contain hot subdwarfs, but are too old, distant and reddened for the blue stragglers to be detectable with UIT. M67 is also sufficiently nearby that all the white dwarfs hotter than $\\sim$ 21,000 K should have been detected on the UIT image. In the population synthesis models of \\markcite{magris93}Magris \\& Bruzual (1993), white dwarfs supply about 10\\% of the 1750~\\AA\\ flux in elliptical galaxies with a weak UV-upturn, and in Section 3.2 we re-examine the white dwarf contribution to galaxy ultraviolet spectra, using more recent stellar atmospheres and evolutionary models. A brief summary of this work was given by \\markcite{last97}Landsman \\& Stecher (1997). A subsequent paper will discuss the UIT observations of the intermediate age ($\\sim$ 2 Gyr) clusters NGC 752 and NGC 7789, which are sufficiently young to allow the turnoff population to be detected in the ultraviolet. ", "conclusions": "There are at least three ways in which Galactic open clusters are an inadequate model for understanding the stellar populations of old galaxies. First, the open clusters have experienced dynamical interactions which can alter the stellar population mix; for example, by leading to evaporation of low-mass stars while concentrating the blue stragglers toward the cluster center. Second, the higher mean stellar density of open clusters might result in enhanced stellar interactions and possibly increased blue straggler or hot subdwarf formation. Finally, due to the relatively small population of open clusters, the rapid evolutionary phases may not be adequately sampled. For example, hot post-AGB stars are believed to be a significant (although not the dominant) contributor to the observed UV-upturn in elliptical galaxies, but consistent with their short ($\\sim 10^5$ yr) lifetime, there are no post-AGB stars present in M67, NGC 188, or NGC 6791. And while hot subdwarfs have now been detected in NGC 188 and NGC 6791, their total number ($<8$) remains uncomfortably low to draw any robust conclusions. Keeping this limitations in mind, we summarize below the main results derived from the UIT observations of M67, NGC 188, and NGC 6791, and their implications for the study of the integrated ultraviolet spectra of old stellar populations. \\begin{enumerate} \\item The UIT image of M67 is dominated by the eleven detected blue stragglers; in particular, the blue straggler F81 contributes 90\\% of the integrated flux of M67 at 1520~\\AA. The ultraviolet flux of the two blue stragglers, F90 and F131, is significantly higher than that predicted on the basis of optical photometry, and probably indicates the presence of hot, subluminous companions. A semi-empirical calculation of the integrated ultraviolet spectrum of M67 shows that, even when neglecting the possibly anomalous star F81, the blue stragglers dominate at wavelengths shorter than $\\sim$ 2600~\\AA. As pointed out by Spinrad et al.\\ (1997), neglect of blue stragglers in population synthesis model fits of the rest-frame near-ultraviolet spectra of a galaxy will result in an underestimate of the galaxy age. \\item Eight white dwarf candidates are identified in the UIT image of M67, including the core-helium white dwarf companion of the yellow giant S1040. Optical spectroscopy of two of these sources has been obtained by \\markcite{flem97}Fleming et al.\\ (1997): G152 is a hot (\\teff $\\sim$ 68,000 K) DA white dwarf, and MMJ 5973 is cooler (\\teff\\ $\\sim$ 18,000 K) DB white dwarf. The number of white dwarf candidates is in reasonable agreement with that expected from theoretical white dwarf cooling models and a cluster age of 4 Gyr. The integrated ultraviolet flux at 1500~\\AA\\ along a white dwarf cooling model is $\\log$ E$_{1500} = -4.23$ L$_{V}^{\\odot}$ Gyr \\AA$^{-1}$, and the contribution of white dwarfs to the integrated spectra of old galaxies is roughly 10\\% of that expected from hot post-AGB stars. \\item Eight probable cluster members are detected on the UIT image of NGC 6791, including the five hot subdwarfs studied spectroscopically by Liebert et al.\\ (1994), and the two additional sdB/O candidates, B9 and B10, reported by \\markcite{kal95}Kaluzny \\& Rucinski (1995). Three probable cluster members are detected on the UIT image of NGC 188, including the sdB spectroscopic binary II-91. The star D702 in NGC 188 is probably a composite including a hot subdwarf, since it has a hot ultraviolet color (\\mbi\\ -- \\mbv\\ = --0.11), but a relatively cool optical color (\\bv\\ = 0.26). The derived luminosity range, $1.10 < \\log$ L/\\lsun\\ $< 1.27$, of the five sdB stars in NGC 6791, and II-91 in NGC 188, is consistent with that expected for metal-rich, hot HB stars. The fraction of hot HB stars in both clusters is about 30\\%, implying that the integrated spectra of both clusters should show a pronounced UV-upturn, as strong as that observed in any elliptical galaxy. \\end{enumerate}" }, "9804/astro-ph9804328_arXiv.txt": { "abstract": "Galaxy distance indicators are subject to different uncertainties and biases which may reflect in the peculiar velocity field. In this work, we present different statistical analysis applied to peculiar velocities of samples of galaxies and mock catalogs taken from numerical simulations considering these observational effects. Our statistical studies comprise the mean relative velocity and velocity correlation for pairs of galaxies as a function of separation, and a bulk flow analysis determined on spheres of $10 h^{-1}$ $Mpc$ radius. In order to test the statistical analysis we use COBE normalized CDM numerical simulations with different density parameters and cosmological constant and take into account the Tully-Fisher (TF) scatter and a possible TF zero-point offset, as well as variations in the results of the simulations arising from different observer positions. We compare the results of the mock catalogs with samples of spiral galaxies taken from the Mark III catalog. The results of our analysis indicate the importance of errors in deriving the density parameter $\\Omega$ from statistics applied to the observational data and it is found that large values $\\Omega \\geq 0.8$ may be derived from the analysis if errors are not taken into account. However, the observed value of the TF scatter ($\\simeq 0.35$ $mag$) make CDM models with $\\Omega > 0.6$, inconsistent with the statistical tests involving relative velocities. A suitable agreement is found for models with $\\Omega \\leq 0.6$, although requiring a TF zero-point offset of the order of a tenth of a magnitude to provide consistency with the observed flow coherence. ", "introduction": "Measurements of peculiar velocities provide direct probes of the mass distribution in the Universe putting constraints on models of large-scale structure formation (e.g. \\cite{peebles}, \\cite{vittorio}). Advances on new distance indicators (\\cite{djor}; \\cite{dresslera}) have allowed estimates of peculiar flows in the local Universe up to $\\sim 50-100 h^{-1}$Mpc (see \\cite{giova}, or \\cite{strauss} for a review). The results from using elliptical galaxies (\\cite{dresslerb}) suggest a large coherence length and amplitude of the local peculiar velocity field. The results from analysis of 1,355 spirals using the Tully-Fisher relation to determine distances (\\cite{mathew}) would lead to an even larger coherence length and a similar amplitude. Other samples are also in agreement with the Great Attractor findings (\\cite{willick}). In a more controversial finding \\cite{postman} find from analysis of clusters of galaxies that the peculiar flows can be coherent over a much larger scale than the Great Attractor findings. All the results indicate significant deviations from the Hubble flow with a large coherence length. Since peculiar velocities probe the mass distribution and also depend on the density parameter $\\Omega$ one can determine the latter by comparing the mass distribution implied by the velocity field with the observed distribution of galaxies. The density parameter is thus determined to within the uncertainty of the bias factor $b$ by determining the factor $\\Omega^{0.6}/b$. Other analysis of the peculiar velocity information may be applied in order to obtain the parameter $\\Omega$. \\cite{berta} developed the POTENT method whereby the mass distribution is reconstructed by using the analog of the Bernoulli equation for irrotational flows. They used the method to analyze to great detail the peculiar velocity field out to about 60$h^{-1}$Mpc (\\cite{bertb}). \\cite{dekel} compared the previously determined velocity field with the observed distribution of galaxies and concluded that the best fit is with $\\Omega^{0.6}/b \\simeq 1$. A similar analysis can be applied to the components of the velocity tensor $U_{ij}$ (\\cite{gorskia}; \\cite{groth}). Recently, \\cite{kashlinsky} has analyzed this tensor for the MarkIII peculiar velocity data to reconstruct the large-scale power spectrum obtaining consistency within the CDM model with $\\sigma_{8} \\Omega^{0.6} \\simeq 0.8$. In a similar study \\cite{zaroubi} obtain a different result $\\sigma_{8} \\Omega^{0.6} \\simeq 0.35$ indicating the need of further analysis. In this paper we study the peculiar velocity field through a statistical analysis of COBE normalized CDM numerical simulations and observational data taken from the Mark III catalog. Our comparison of models and observations take into account observational uncertainties and possible biases, and variations of the results according to different observers in fully non-linear numerical simulations. \\section {Data} We use the Mark III catalog (\\cite{markIII1};\\cite{markIII2};\\cite{markIII3}) as a suitable data set to analyze the peculiar velocity flow. This Catalog lists Tully-Fisher and $Dn-\\sigma$ distances and radial velocities for spiral, irregular, and elliptical galaxies. In our analysis we have used only spiral galaxies given their large number and smooth spatial distribution (see Table 1). \\begin{deluxetable}{cccc} \\tablewidth{35pc} \\tablecaption{Observations: The Mark III spirals} \\tablehead{ \\colhead{Subsample}& \\colhead{$N^o$ of Gx.}& \\colhead{TF relation}& \\colhead{$\\sigma_{TF}$} } \\startdata \\phm{II} Aaronson et al. Field & 359 & $M_H=-5.95+10.29 \\eta$ & $0.47$ \\nl \\phm{II} Mathewson et al. (1992) & 1355 & $M_I=-5.79+6.8 \\eta$ & $0.43$ \\nl \\phm{II} Willick, Perseus Pisces (1991) & 383 & $M_r=-4.28+7.12 \\eta$ & 0.38 \\nl \\phm{II} Willick, Cluster Galaxy (1991) & 156 & $M_r=-4.18+7.73 \\eta$ & $0.38$ \\nl \\phm{II} Courteau-Faber (1993) & 326 & $M_r=-4.22+7.73 \\eta$ & $0.38$\\nl \\phm{II} Han-Mould et al., Cl. Gx. (1992) & 433 & $M_I=-5.48+7.87 \\eta$ & 0.4 \\nl \\enddata \\end{deluxetable} The velocity parameter $\\eta = Log \\Delta V - 2.5$ is determined either from HI profiles or from optical $H_{\\alpha}$ rotation curves. The TF relations and their corresponding scatters for the different samples of spiral galaxies are given by \\cite{markIII3} and are shown in Table 1, where the absolute magnitud $M$ satisfies $M=m-5\\log cz$. The galaxy apparent magnitudes $m$ of the Tully-Fisher distances are corrected for Galactic extinction, inclination and redshift (see Willick et al. 1997 for details). The selection bias in the calibration of the forward TF relation can be corrected once the selection function is known. But then, the TF inferred distances and the mean peculiar velocities suffer from Malmquist bias. Suitable procedures to consider these biases, induced both by inhomogeneities and selection function, have been discussed (see for instance /cite{freud95} and references therein) where the spatial distribution, selection effects and observational uncertainties are realistically modeled through Monte-Carlo simulations. We have used in our analysis forward TF distances, fully corrected for Inhomogeneous Malmquist Bias (\\cite{markIII1}, \\cite{markIII2}, \\cite{markIII3}). Nevertheless we have found that the results of the statistics studied in this work do not change significantly if inverse TF distances are used as it will be discussed below. Radial velocities used to infer the peculiar velocity of the galaxies are referred to the Cosmic Microwave Background frame. It should be remarked that galaxy distance estimates are subject to errors due to the scatter in the TF relation (\\cite{mo}; \\cite{willickth}; \\cite{mathew}) and uncertainties of the TF zero-point (\\cite{shanks}; \\cite{willickth}). Also, the possible presence of a small fraction of spurious velocities in the data induced by either galaxy peculiarities or observational errors in distance estimates (\\cite{jacoby}) should be taken into account. ", "conclusions": "In this work we have attempted to asses the effects of galaxy distance uncertainties and biases on statistical tests of the peculiar velocity field using samples of spiral galaxies from the Mark III compilation and mock catalogs taken from numerical simulations. Our studies comprise relative velocity tests and pair velocity correlations as a function of separation, as well as a bulk flow analysis determined on spheres of $10 h^{-1}$ $Mpc$ radius. We construct mock catalogs using numerical simulations corresponding to COBE normalized CDM models with different values of density parameter and cosmological constant. The models take into account the Tully-Fisher (TF) scatter and possible TF zero-point shift, as well as variations in the results of the simulations arising from different observer positions. The analysis of the departures from gaussianity of the observational velocity distributions as measured by the kurtosis tests $K$ and $K1$ show that only a small fraction ($<2 \\%)$ of possible spuriously high velocities (and therefore biased distance estimates) might be present in the observational data. We find that the observed scatter of the TF relation plays a very important role when deriving constraints to the cosmological parameters. If errors in galaxy distance estimates are neglected, the observed magnitude of the $\\Delta V$ results show consistency with high density CDM models ($\\Omega>0.8$). When the observed TF scatter $\\Delta M \\simeq 0.35$ is taken into account a significant disagreement with observations is found for the $\\Delta V$ and $D1$ statistics in models with $\\Omega > 0.6$. A possible offset in the TF zero-point used to determine the distances of the galaxies in the observational data artificially enhance velocity correlations ($\\Pi$ and $M$ statistics). Due to this fact, if zero-point offsets of the order of a tenth of a magnitude are considered, the CDM models explored provide a more satisfactory fit to observations in these tests. The results presented in this work are not sensitive to sample variations. We have applied the same statistical tests to the subsample corresponding to Mathewson data and we find similar results than those of the total sample. It should be noted that the variations on the statistics arising from different observers are significantly enhanced when errors on the peculiar velocities are included. This fact make more difficult the distinction between the models. Error bars in figures 1 to 5 show the variations of the results arising from different observer positions in the models and serve as a test for the dependence of the observational results on our particular position in space. In our analysis of biased models we have not found relevant differences in the results of the statistical analysis of samples with different density thresholds. The selection of low (high) density particles in the simulations does not produce very relevant effects in the statistics although lower (higher) values of $\\Pi$ and $\\Delta V$ are observed due to the oversampling (undersampling) of high density particles where random motions dominate. Finally when comparing models and observations, we find that neither the K, K1, M, nor $\\Pi$ statistical tests are sensitive to the density parameter of the models. We find crucial to consider properly the intrinsic scatter of the Tully-Fisher relation in studies of the peculiar velocity field. Although low values of the density parameter $\\Omega < 0.6 $, are favored by our statistical analysis $D1$ and $\\Delta V$ when this scatter is taken into account $\\pm 0.15$ $mag$ zero-point offsets in the TF relation are required in order to provide the observed values of $\\Pi$ and $M$ statistics in the models explored. A positive TF zero-point offset of this magnitude would imply lower values of the Hubble constant which may be worth to consider in a controversial topic (see for instance \\cite{shanks} and \\cite{th} and references therein)." }, "9804/astro-ph9804091_arXiv.txt": { "abstract": "The conversion of neutron matter into strange matter in a neutron star occurs through the non-leptonic weak-interaction process. We study the energy loss of the neutron star by the emission of axions in that process. Owing to that process, the neutron star will liberate the energy which can in no way be negligible as an axion burst. \\\\ \\\\ PACS numbers : 14.80.Mz, 95.30.Cq, 97.60.Jd \\\\ Keywords : axions, astrophysics, neutron stars, strange star \\\\ \\\\ ", "introduction": "The possibility of the existence of stable quark matter in the early universe or inside a neutron star or in relativistic heavy-ion collision experiments has been studied in many works \\cite{sqm}\\@. In this work, we consider the conversion of a neutron star into a strange star and its energy loss during the conversion process. In order to form strange matter in the interior of a neutron star, neutron matter should be converted into strange matter. Several conversion mechanisms have been discussed by Alcock et al.\\ \\cite{alcock}\\@. When a neutron and a stable strangelet (strange quark matter droplet) meet, the neutron is readily absorbed, while a proton can coexist with a strangelet due to the Coulomb barrier between them \\cite{alcock}\\@. Therefore, a stable strangelet as a seed for the process will trigger a conversion in such a way that it grows by absorbing neutrons and, eventually, convert most of the neutron star into a strange star \\cite{olinto,daietal}\\@. Then the conversion would liberate about $10^{52}$ ergs in binding energy \\cite{olinto}\\@. A variety of mechanisms have been suggested for seeding the interior of a neutron star with stable strangelets so far \\cite{alcock,lugones}\\@. They can be divided into two main categories \\cite{olinto}; (1) the primary mechanisms in which the seed is formed inside the neutron star, (2) the secondary mechanisms in which the seed comes from the interstellar medium. In a viewpoint of the primary mechanisms \\cite{alcock}, the central high densities and pressures in a neutron star may be sufficient for a phase transition from neutron matter to two-flavor quark matter to occur. Subsequently two-flavor quark matter can then easily decay into the lower energy strange matter through weak interactions. If we assume that strangelet has formed inside a neutron star, subsequent conversion process of the rest of the star is described as follows \\cite{olinto,olesen}\\@. The volume over which strange matter equilibrates was shown to be much smaller than that of the total strange matter region \\cite{lugones,horvath}, so that the problem can be treated hydrodynamically in one-dimensional geometry. As the conversion front sweeps into neutron matter, the small region behind the conversion front has an excess of down quarks relative to strange quarks due to the flux of neutrons ($udd$) at the conversion front. The excess down quarks will convert into strange quarks via the non-leptonic weak process \\cite{madsen}, $d \\; + \\; u \\; \\rightarrow \\; s \\; + \\; u$, as long as $\\mu_d > \\mu_s (n_d > n_s)$\\@. By this process a $d$ quark can change itself to an $s$ quark until the Fermi energies of all the flavors become equal and the energy per baryon drops in comparison to the ordinary two-component nuclear matter. Other leptonic decay processes are suppressed considerably \\cite{madsen}\\@. The conversion will liberate $\\sim 10^{52}$ ergs of energy (assuming that $\\sim$ 10 $MeV$ is liberated per neutron converted)\\@. This energy will be radiated as neutrinos, photons, $e^{+}e^{-}$ pairs, and axions, etc. In this stage neutrinos and axions can escape the star. Many authors \\cite{daietal,aggs} have considered the escape of neutrinos. In this work we calculate the energy loss due to the emission of axions in the process of such a conversion of a neutron star into a strange star. At the chemical equilibrium between the quarks and the electrons, non- or semi-leptonic weak interaction is not important in neutrino and axion emission in quark matter because the weak interactions coupling $s$ and $u$ quarks are Cabibbo suppressed relative to the interactions coupling $d$ and $u$ quarks. For the nonequilibrium processes such as the conversion from two-flavor to three-flavor quark matter, however, as the strange quark semi-leptonic processes in neutrino emissions and non-leptonic processes in axion emissions are of significance. The axion is a pseudo-Goldstone boson which was introduced by Peccei and Quinn (PQ) to solve the strong CP-problem in a natural way \\cite{PQ}\\@. However, theoretical and experimental investigations give little guidance on the PQ symmetry-breaking scale, $f_a$, and therefore on the mass of the axion \\cite{WW}\\@. The axion decay constant $f_a$ is related to the axion mass \\cite{axion} \\[ m_a \\simeq \\left( \\frac{0.62 \\times 10^{7} GeV}{f_a} \\right) \\; eV . \\] There are two generic types of invisible axions; the Dine-Fischler-Srednicki-Zhitnitskii (DFSZ) axion which couples to both quarks and leptons at tree level \\cite{DFSZ} and the hadronic, or Kim-Shifman-Vainshtein-Zakharov (KSVZ), axion which has no tree-level coupling to leptons but does couple to them at loop levels \\cite{KSVZ}\\@. In invisible axion models, the axion mass $m_a$ is in principle arbitrary, however astrophysical and cosmological considerations \\cite{axion} can provide an upper and lower bounds for $m_a$\\@. The astrophysical bounds on $m_a$ are due to the fact that axion emission is an additional energy loss mechanism for stars. If such axions could be copiously produced during the conversion of a neutron star into a strange star, it might drastically alter the energy budget of stars. The axions might thus deplete the stellar energy and change the usual course of stellar evolution. The emission of axions would have hastened the cooling process. The quarks are expected to be highly relativistic and degenerate in such matter with their Fermi energies in the range of 300 to 500 $MeV$\\@. A calculation of the axion energy flux from strange quark matter has been made by Anand et al.\\ \\cite{anand}\\@. They have shown that the axion emission rate is several orders of magnitude smaller compare to the neutrino emission rate. We investigate here the energy loss of the neutron star during the conversion of non-strange quark matter into strange quark matter in the interior of the star. We assume that the axions escape freely out of the star as soon as they are produced, and that the temperature in the interior is constant during the conversion for the reasons that strong and electromagnetic interaction timescales are much smaller than those of weak interactions. Unless otherwise noted all equations assume $\\hbar = c = k_B = 1$\\@. ", "conclusions": "Our main result for the total energy flux rate for the emission of axions from conversion of non-strange quark matter into strange quark matter is given by Eq. (12). Here we can carefully quote our bound on the axion coupling $f_a$, using the simple inequality \\cite{ishizuka}, \\be {\\cal E}_a \\cdot V \\cdot \\delta t < E_m \\ee The allowed energy loss, $E_m / \\delta t$ is taken to be $10^{52}$ erg/sec for definiteness, and the volume of axion emitting region $V = \\frac{4}{3} \\pi R^3$ with $R = 10 \\sim 20 km$\\@. As for the conversion timescale, if $T=10 MeV$, it varies roughly between one and ten minutes \\cite{olinto}\\@. So, we take the timescale typically $\\delta t = 100 sec$\\@. Then, from Eq. (14), we can obtain the bound on $f_a$, \\be \\sim 10^5 \\; GeV < f_a \\,. \\ee We note that the bound based on Eq. (14) is legitimate only if the emitted axions never interact on their way out of the star. If the axion coupling is large enough, axions once produced may interact many times, namely be absorbed and reemitted more than once in the hot core of the star, and may be trapped thermally. As applications, Mikheev and Vassilevskaya \\cite{mik} recently investigated the radiative decay of the axion $a \\rightarrow \\gamma\\gamma$ in an external electromagnetic field in the DFSZ model. They concluded that the decay probability is strongly catalyzed by the external field, namely, the field removes the main suppression caused by the smallness of the axion mass. Therefore, the radiative decay of the axion in strong magnetic fields of the neutron star could give interesting astrophysical phenomena. Furthermore, we can consider that the sudden conversion from a neutron star to a strange star may account for the gamma ray bursts at comological distances \\cite{olinto}\\@. Most of the energy is probably released in the form of neutrinos. If a part of the total energy goes into $\\gamma$ rays decayed from the axions, it will be large enough to account for the gamma ray bursts at cosmological distances and to explain their isotropic distribution. The outcome of such a conversion event will be the emission of as much as $\\sim 10^{58} MeV$ of energy in $\\sim sec$ to $\\sim min$ as radiation with a typical temperature of tens of $MeV$\\@. This conversion can be observed as a gamma ray burst \\cite{olinto}\\@. In summary, as a possible mechanism for production of axions, we considered the conversion of non-strange quark matter into strange quark matter. We also estimated the energy loss of neutron stars through the emission of axions in addition to the cooling provided by the neutrino emission. During the conversion period, the important process is the non-leptonic weak-interaction. We assume that the axions are not trapped and escape freely out of the star as soon as they are produced, and that the temperature in the interior is constant during the conversion. We find that the energy carried away by axions during the conversion is not negligible. It is also found that the axion emission rate is three orders of magnitude larger compared to the chemical equilibrium case. In addition, we discussed the bound on the axion coupling. \\vspace*{3cm}" }, "9804/astro-ph9804058_arXiv.txt": { "abstract": "We have analyzed deep $B$ and $V$ photometry of the Carina dwarf spheroidal reaching below the old main-sequence turnoff to $V \\sim 25$. Using simulated color-magnitude diagrams to model a range of star formation scenarios, we have extracted a detailed, global star formation history. Carina experienced three significant episodes of star formation at $\\sim 15$ Gyr, 7 Gyr, and 3 Gyr. Contrary to the generic picture of galaxy evolution, however, the bulk of star formation, at least 50\\%, occured during the episode 7 Gyr ago, which may have lasted as long as 2 Gyr. For unknown reasons, Carina formed only 10-20\\% of its stars at an ancient epoch and then remained quiescent for more than 4 Gyr. The remainder ($\\sim 30\\%$) formed relatively recently, only 3 Gyr ago. Interest in the local population of dwarf galaxies has increased lately due to their potential importance in the understanding of faint galaxy counts. We surmise that objects like Carina, which exhibits the most extreme episodic behavior of any of the dwarf spheroidal companions to the Galaxy, are capable of contributing to the observed excess of blue galaxies at $B \\sim 24$ only if the star formation occurred instantaneously. ", "introduction": "The Carina dwarf galaxy, one of the nine known dwarf spheroidal companions to the Galaxy, has a surprisingly complex star formation history. Detection of carbon stars provided the first suggestion of a significant intermediate-age population (Cannon, Niss, \\& Norgard-Nielsen 1981), followed by main-sequence photometry which revealed a young turn-off due to a population perhaps only 6--9 Gyr old (Mould \\& Aaronson 1983; MA hereafter). MA fit a simulated luminosity function comprised of 7 Gyr-old stars to their data, and estimated the contribution of an old population to be relatively small in comparison. Saha, Monet, and Seitzer (1986) observed a large number of RR Lyraes in the central region of the galaxy and established a lower limit of 2--3\\% for the fraction of old stars, depending on the yield of RR Lyraes. Mighell (1990) estimated the relative sizes of the populations from the double-peaked color distribution near the MSTO region to be 85\\% for the intermediate-age burst and 15\\% for the old episode. Using simulated luminosity functions, Mighell \\& Butcher (1992) later fit intermediate-age burst models to this deep, main-sequence photometry and estimated an upper limit to the old population of 40\\%. More recently, Smecker-Hane, et al. (1994; hereafter SHSHL) resolved two separate horizontal branches (HBs). Although the separation is not independent of metallicity effects, it is very suggestive of the multi-episode nature of Carina's history. We are left with an estimate between 2\\% and 40\\% for the fraction of old stars, and no clear result on the duration of the star formation episodes. The chemical evolution of the galaxy is likewise still in question. Spectroscopy of 15 giants in Carina (Da Costa 1994) resulted in an average [Fe/H] of --1.9, with one giant significantly more metal poor ([Fe/H] = --2.2). The excellent areal coverage of the SHSHL data reveal a very thin giant branch. Given the huge apparent age spread, a metallicity spread may be necessary to compensate and produce the observed narrow RGB. Improved methods of analysis are needed. Detailed analyses of stellar luminosity functions have some advantages over more traditional isochrone fitting. These methods, however, were designed for coeval systems. Because dwarfs exhibit a range of ages and metallicities, these approaches are problematic at best, and misleading at worst. A much better way to unravel complex star formation histories like those exhibited by dwarf galaxies such as Carina and Leo~I is to use all of the information embodied by color-magnitude diagrams. In these cases, where the galaxies appear to have experienced bursts at intermediate epochs, a conventional luminosity function would not distinguish between the old subgiant branch and the young MS stars, for example. Color information proves essential to resolve multiple components in these systems. Simulated color-magnitude diagrams have been used successfully in studies of bright stars in dwarf irregulars (Tosi et al. 1991; Tolstoy 1996; Gallart et al. 1996a, 1996b, Aparicio et al. 1997a, 1997b), of LMC field stars (Bertelli et al. 1992; Vallenari et al. 1996), and of LMC clusters (Vallenari et al. 1994). We have deep CCD photometry (limiting magnitude $V = 24.5$) with reasonably small photometric errors ($\\sim 0.02$ at $V=23$) of three fields in the Carina dwarf galaxy, reaching well below the old main-sequence turnoff. Because the photometry reaches below the old MSTO, these data are well-suited to a detailed extraction of the star formation history by comparing model color-magnitude diagrams to that of Carina. Section 2 is a discussion of the observations and reductions of the Carina data. Section 3 outlines the analysis applied to these data, including the development of the pseudo-LF and the generation of similulated color-magnitude diagrams. The results, presented in section 4, are summarized and briefly discussed in the context of galaxy evolution in section 5. ", "conclusions": "Carina experienced an episode of star formation 7 Gyr ago which lasted no more than 2 Gyr and which was responsible for at least 50\\% of its stars. The old population ($12-15$ Gyr) may amount to $10\\%-20\\%$ of Carina. The bulk of the remainder (($20\\%-30\\%$) is relatively young, between 2.5 and 3.5 Gyr old. While the details of one galaxy's SF history may seem inconsequential, taken in context and as a member of the Local Group dwarf population, the details are relevant to deeper, unresolved issues in cosmology and galaxy evolution. In the case of Carina, all studies confirm that at least one relatively long pause in star formation lasting $\\sim 4$ Gyr occurred. The mechanism by which this kind of low mass system could experience such a pause and then another strong burst of star formation is not understood. Current ISM simulations can produce such a large gap in an isolated dwarf only by ejecting the gas out to $\\sim 20$ kpc (Babul 1996). A dwarf such as Carina, residing in the potential well of the Galaxy, should lose that gas, preventing any further episodes of star formation. Recent Hubble Space Telescope observations of Carina have been studied by Mighell (1997). He claims to detect a significant number of stars in the region of the gap in star formation lasting from roughly 8 to 12 Gyr ago. This is interpreted as indicative of significant star formation which began in the central region of Carina, and propagated to the outer regions. However, the number of stars in the same section of our CMD, derived from fields several arcminutes from the center of Carina and overlapping the Mighell (1990) field, is consistent with the number in the HST CMD. Whether or not the number of stars in this area implies significant on-going star formation, there does not seem to be evidence of a difference between the central region and regions further out, such as that seen in the recently-discovered Antlia dwarf galaxy (Aparicio et al 1997c). Assigning ages to individual faint stars based on the isochrones which constrain them in color and magnitude must be handled with care. Firstly, at faint magnitudes, stars have a larger photometric error and thus a larger implied age range. Age determinations for specific stars are therefore subject to greater statistical uncertainty. Secondly, the lifetime of stars crossing the Hertzprung gap increases as stars become less massive. This must be accounted for when predicting the relative contributions of different ages. In light of these considerations, the results of the WFPC2 study may not differ from earlier ground based results. Evidence of intermediate-age stars in the halo leads to the question of whether dwarf spheroidals shredded by the Galaxy could be a significant source of the halo population (Preston et al. 1994; Mateo 1996). The Sagittarius dSph is currently being ripped apart by the Milky Way, proving that this type of interaction between a large galaxy and its tiny neighbors does occur (Ibata, Gilmore, \\& Irwin 1994; Mateo et al. 1995). In addition, the dSph, especially Carina and Leo I, contain significant numbers of intermediate-age stars. Mateo (1996) shows that the fraction of relatively young and intermediate-age stars in the entire population of dwarf spheroidals is not inconsistent with the fraction in the Galactic halo. The detailed interpretation of deep galaxy counts and redshift surveys depends on the SFH of the dwarfs which comprise the excess of faint galaxies. For example, `flashing' dwarfs - i.e. bursting quickly and intensely - contribute to the deep counts to a dramatically different extent than do dwarfs that experience more drawn-out bursts that slowly turn on and off (Campos 1997). The prevalence of bursting behavior in the SFH of Local Group dwarfs may statistically constrain the degree to which galaxies similar to these local dwarfs could be responsible for the faint excess. We can place an upper limit on the star formation rate (SFR) of Carina 7 Gyr ago by assuming that the episode of star formation took place instantaneously. By choosing a reasonable model, we can easily calculate the total number of stars formed during this episode in the volume of the galaxy covered by our three fields, taking into account the incompleteness. We can use this number to normalize the IMF, and then integrate to get the total mass formed during this episode. The SFR is then given by: \\begin{displaymath} SFR = \\frac{M_{total}}{\\Delta t}*\\frac{1}{\\gamma } M_{\\odot } yr^{-1} \\end{displaymath} where $\\Delta t$ is the duration of the episode in years and $\\gamma$ is the fraction of the galaxy's total surface brightness in the three fields. If the episode lasted 10 Myr, the implied SFR for Carina at that time would be roughly 0.8 $M_{\\odot } yr^{-1}$. If the episode lasted 2 Gyr, the SFR is down by more than a factor of 100 to 0.004 $M_{\\odot } yr^{-1}$. Either SFR is comparable to the instantaneous SFR in even such luminous objects as blue compact dwarfs (Fanelli et al. 1988). Could Carina-type galaxies contribute to the excess of faint galaxies at around $B \\sim 24$? These galaxies have been shown to have redshifts between 0.3 and 0.7 (Glazebrook et al. 1995). Assuming $H_{\\circ}$ = 50 km/s/Mpc, and $\\Omega$ = 1, 7 Gyr corresponds to $z \\sim 0.5$, which places the episode of star formation in the appropriate epoch. If the burst were essentially instantaneous, the implied total luminosity of Carina at 7 Gyr would be $\\sim 2 \\times 10^8 L_\\odot$. Carina would be visible to $z \\sim 0.3$ in a sample limited to $V \\sim 25$, putting it at the near edge of objects that could contribute to these counts. Galaxies 3-10 times more luminous would fall in the range $z \\sim 0.5$ to 1.0. Although the local dSph cover a range up to 50 times the luminosity of Carina, Carina is the most extreme example of this type of episodic behavior. Further, if the episodes of star formation are extended in time, then these dwarfs would only be visible locally. Thus, despite concerted efforts at unravelling the SFH of these galaxies, their role in faint galaxy counts problem remains unresolved." }, "9804/astro-ph9804172.txt": { "abstract": "We have analyzed the projected galaxy distributions in a subset of the ENACS cluster sample, viz. in those 77 clusters that have z $<$ 0.1 and R$_{\\rm ACO} \\ge$ 1 and for which ENACS and COSMOS data are available. For 20~\\% of these, the distribution of galaxies in the COSMOS catalogue does not allow a reliable centre position to be determined. For the other 62 clusters, we first determined the centre and elongation of the galaxy distribution. Subsequently, we made Maximum-Likelihood fits to the distribution of COSMOS galaxies for 4 theoretical profiles, two with `cores' (generalized King- and Hubble-profiles) and two with `cusps' (generalized Navarro, Frenk and White, or NFW, and de~Vaucouleurs profiles). We obtain average core radii (or characteristic radii for the profiles without core) of 128, 189, 292 and 1582 kpc for fits with King, Hubble, NFW and de~Vaucouleurs profiles respectively, with dispersions around these average values of 88, 116, 191 and 771 kpc. The surface density of background galaxies is about 4 10$^{-5}$ gals arcsec$^{-2}$ (with a spread of about 2 10$^{-5}$), and there is very good agreement between the values found for the 4 profiles. There is also very good agreement on the outer logarithmic slope of the projected galaxy distribution, which is that for the non-generalized King- and Hubble-profile (i.e. $\\beta_{King}$ = $\\beta_{Hubble}$ = 1, with the corresponding values for the two other model-profiles). We use the Likelihood ratio to investigate whether the observations are significantly better described by profiles with cusps or by profiles with cores. Taking the King and NFW profiles as `model' of either class, we find that about 75 \\% of the clusters are better fit by the King profile than by the NFW profile. However, for the individual clusters the preference for the King profile is rarely significant at a confidence level of more than 90 \\%. When we limit ourselves to the central regions it appears that the signifance increases drastically, with 65 \\% of the clusters showing a strong preference for a King over an NFW profile. At the same time, about 10 \\% of the clusters are clearly better fitted by an NFW profile than by a King profile in their centres. We constructed composite clusters from the COSMOS and ENACS data, taking special care to avoid the creation of artificial cusps (due to ellipticity), and the destruction of real cusps (due to non-perfect centering). When adding the galaxy distributions to produce a composite cluster, we either applied no scaling of the projected distances, scaling with the core radii of the individual clusters or scaling with r$_{200}$, which is designed to take differences in mass into account. In all three cases we find that the King profile is clearly preferred (at more than 95 \\% confidence) over the NFW profile (over the entire aperture of 5 core-radii). However, this `preference' is not shared by the brightest (M$_{b_j}$ $\\la$ -18.4) galaxies. We conclude that the brighter galaxies are represented almost equally well by King and NFW profiles, but that the distribution of the fainter galaxies clearly shows a core rather than a cusp. Finally, we compared the outer slope of the galaxy distributions in our clusters with results for model calculations for various choices of fluctuation spectrum and cosmological parameters. We conclude that the observed profile slope indicates a low value for $\\Omega_0$. This is consistent with the direct estimate of $\\Omega_0$ based on the $\\frac ML$-ratios of the individual clusters. ", "introduction": "Until fairly recently, the projected galaxy density in rich galaxy clusters was generally described by King or Hubble profiles. In these profiles, the logarithmic slope of the mass distribution is essentially zero near the cluster centre. The core radius which is the characteristic scale of the distribution, was sometimes also regarded as the distance which more or less separates dynamically distinct regions in a cluster. From the kinematics of the galaxy population it appears that in clusters the relaxation time is significantly shorter than the Hubble time {\\em only} in the very central region within at most a few core radii (see e.g. den Hartog and Katgert 1996). The concept of cores in clusters has been seriously challenged, on observational grounds (e.g. Beers $\\&$ Tonry 1986) and as a result of numerical simulations. Navarro, Frenk and White (1995, 1996) found e.g. that the equilibrium density profiles of dark matter halos in universes with dominant hierarchical clustering all have the same shape, which is essentially independent of the mass of the halo, the spectrum of initial density fluctuations, or the values of the cosmological parameters. This `universal' density profile (NFW profile hereafter) does not have a core, but has a logarithmic slope of --1 near the centre which, at large radii, steepens to --3, and thus closely resembles the Hernquist (1990) profile except for the steeper slope of the latter at large radii of --4. Navarro, Frenk and White (1997) argue that the apparent variations in profile shape, as reported before, can be understood as being due to differences in the characteristic density (or mass) of the halo, which sets the linear scale at which the transition of the flat central slope to the steep outer slope occurs. They also argued that the existence of giant arcs in clusters requires that the mass distributions in clusters does not exhibit a flat core in the centre. In other words: if clusters have cores, the lensing results require that the core radii are very small, at least quite a bit smaller than the values usually quoted. It is not clear that galaxy clusters should have cores; after all, the dynamical structure of galaxy clusters is quite different from that of globular clusters, for which Michie \\& Bodenheimer (1963) and King first proposed density profiles with cores, in particular the King profile (see e.g. King 1962). On the other hand, the X-ray data for clusters are quite consistent with the existence of a core in the density distribution. More specifically, it was argued recently by Hughes (1997) that the NFW profile would induce a temperature gradient. The existence of such a gradient in the Coma cluster can be excluded at the 99$\\%$ confidence level. Similarly, the galaxy surface density in clusters is generally found to be consistent with a King profile. For galaxy clusters, little use has been made of the de Vaucouleurs profile to describe the galaxy density, even though the latter was found to arise quite naturally in N-body simulations of the collapse of isolated galaxy systems (e.g. van Albada 1982). In view of the claimed universality of the NFW profile found in the simulations, it seems useful to have a closer look at the projected distribution of the galaxies in clusters. After all, the NFW profile refers to the total gravitating mass, and it is not obvious that the galaxy distribution should have exactly the same shape as the distribution of total mass; although in numerical experiments no strong biasing between dark and luminous matter in clusters was seen (e.g. van Kampen 1995). In this respect, it is noteworthy that Carlberg et al. (1997) find that the combined galaxy density profile of 16 high-luminosity X-ray clusters at a redshift of $\\approx$0.3 closely follows the NFW profile. More precisely, the logarithmic slope in the central region is consistent with the value of --1 of the NFW and Hernquist profiles, while the outer slope is consistent with both --3 (the NFW value) and --4 (the Hernquist value). The outer slope of the density profile was found by several authors (e.g. Crone et al. 1994, Jing et al. 1995, and Walter and Klypin 1996) to reflect the details of the formation scenario, and in particular the value of the density parameter of the universe. In addition, this slope is unlikely to be constant in time but is expected to get steeper with decreasing redshift. For that reason, it is important to study both the characteristics of the density profiles of rich clusters and their dependence on redshift. In this paper, we investigate the projected galaxy distributions for a sample of 62 rich and nearby (z $\\la$ 0.1) clusters. These clusters are taken from the volume-limited ENACS (ESO Nearby Abell Cluster Survey) sample of R$_{\\rm ACO} \\ge 1$ clusters (see e.g. Katgert et al. 1996 (paper I), Mazure et al. 1996 (paper II), Biviano et al. 1997 (paper III), Adami et al. 1998 (paper IV), Katgert et al. 1998 (paper V) and de Theije $\\&$ Katgert 1998 (paper VI)). In $\\S$ 2, we first describe the sample of clusters that we used, and the data on which we based our analysis. In $\\S$ 3, we discuss the results of Maximum-Likelihood fitting of profiles with and without a core, to the individual galaxy distributions taken from the COSMOS catalogue. In $\\S$ 4 we discuss the galaxy density distribution for composite clusters (COSMOS and ENACS), in which the individual clusters are combined. In $\\S$ 5 we discuss the constraints provided by the outer slope of the density distributions for the parameters of the formation scenario and in $\\S$ 6 we present the conclusions. ", "conclusions": "We have studied the projected galaxy distributions in 77 clusters from the ESO Nearby Abell Cluster Survey. The present sample is an unbiased subset of the volume-limited ENACS sample, and thus forms a representative local (z $<$ 0.1) sample of rich (R$_{ACO}$ $>$ 1), optically selected clusters. We used both COSMOS and ENACS data to test the character of the projected galaxy distributions. In particular, we have investigated whether the galaxy distributions in rich clusters have cusps or cores in their central regions. We have made maximum Likelihood fits to the observed distribution of COSMOS galaxies to solve for the position and the elongation of the clusters. For 15 of the 77 clusters, no reliable centre could be determined and these clusters were not considered further. Using the positions and elongations, we subsequently solved for each of the 62 remaining clusters the three parameters that describe each of the four theoretical profiles that we tested, as well as the density of background galaxies. The four model profiles that we tested against the data are the King, Hubble, NFW (Navarro, Frenk and White) and the de~Vaucouleurs profiles. Although the solutions do not converge for all of the clusters nor for all four profiles, we obtain reliable results for between 75 and 95 \\% of the clusters (depending on the model profile). We find mean values for r$_c$, the characteristic scale of the 2-D galaxy distribution, and dispersions around the means of 128 $\\pm$ 88, 189 $\\pm$ 116, 292 $\\pm$ 191 and 1582 $\\pm$ 771 kpc, for the King, Hubble, NFW and de Vaucouleurs profiles respectively. The outer logarithmic slopes of the distributions were generalized by the usual $\\beta$-parameter, which we find to have the following average values: 1.02 $\\pm$ 0.08, 1.03 $\\pm$ 0.07, 0.61 $\\pm$ 0.05 and 7.6 $\\pm$ 0.5, for the King, Hubble, NFW and de Vaucouleurs profiles respectively, which are consistent. The average background density at the limit of the COSMOS catalogue is about 4 10$^{-5}$ galaxies arcsec$^{-2}$. In order to investigate whether the galaxy distributions in our clusters preponderantly have cores or cusps, we have determined the likelihood ratio for the King and NFW profiles. Using all galaxies down to the COSMOS magnitude limit of about m$_{b_j}$ $\\approx$ 19.5, we find that in general the King profile is more likely to be a good representation of the data than the NFW profile. However, for the individual clusters this preference for the King profile is generally not statistically significant. If we restrict the analysis to the central regions, the significance of the preference for the King-profile fits increases, even though the number of galaxies decreases. We have increased the statistical weight for the likelihood analysis by combining the galaxy distributions in a subset of 29 of the 62 clusters, which show a regular galaxy distribution. We take special care to avoid the creation of an artificial cusp (by taking the ellipticities into account), and to avoid the destruction of a real cusp by summing distributions with different scale lengths. We have also checked that it is unlikely that the uncertainty in the centre positions has erased a cusp. For the test we summed without scaling projected distances, after scaling with r$_c$, as well as with r$_{200}$. In all three cases we find that the King profile provides a better fit to the data than the NFW profile, at confidence levels of more than 95 \\%. Interestingly, this preference is not shared by the brighter galaxies. Finally, we have used the outer profile slope (i.e. the result that $\\beta$ is very close to 1.0), in combination with several results from numerical models to conclude that the density parameter $\\Omega _0$ is likely to be considerably smaller than unity. In addition, the available models indicate that the Universe probably has an open geometry (i.e. no closure through $\\Lambda $ is indicated). This low implied value of $\\Omega _0$ is fully consistent with a direct determination based on the $\\frac ML$ ratios of our clusters." }, "9804/astro-ph9804085_arXiv.txt": { "abstract": "We perform a fractal analysis of the Southern Sky Redshift Survey 2, following the methods prescribed by Pietronero and collaborators, to check their claim that the galaxy distribution is fractal at all scales, and we discuss and explain the reasons of some controversial points, through tests on both galaxy samples and simulations. We confirm that the amplitude of the two--point correlation function does not depend on the sample depth, but increases with luminosity. We find that there is no contradiction between the results of standard and non--standard statistical analysis; moreover, such results are consistent with theoretical predictions derived from standard CDM models of galaxy formation, and with the hypothesis of homogeneity at large scale ($\\sim 100$ \\h). However, for our SSRS2 volume--limited subsamples we show that the first zero--point of the autocorrelation function $\\xi(s)$ increases linearly with the sample depth, and that its value is comparable to the radius of the maximum sphere which can be completely included in the sampled volume; this implies that the true zero--crossing point of $\\xi(r)$ has not been reached. We conclude that the apparent fractal behavior is due to a combination of a luminosity--dependent correlation amplitude and the recovering of power at larger scales in deeper samples. ", "introduction": "One of the pillars of the standard cosmological models is the large--scale homogeneity of the Universe (e.g. Peebles 1993). The standard statistical methods to analyze the large--scale structure, as the two--point correlation function $\\xi(r)$, are described by Peebles (1980); they rely on the definition of a mean galaxy density $\\bar{n}$, which is meaningful only if the assumption of large--scale homogeneity is true. However, at small scales the autocorrelation function of the galaxy distribution is positive and can be fitted by a power--law, which is a property of a fractal set (see Peebles 1980). On the basis of the observational evidence that galaxies are clustered in ever increasing systems, from groups and clusters to superclusters, de Vaucouleurs (1970, 1971) presented ``the case for a hierarchical cosmology''\\footnote{The idea of a hierarchical Universe has a long history, which dates back to the XVIII$^{th}$ century (see E. Harrison, 1987, {\\em Darkness at night, A Riddle of the Universe}, Harvard University Press, Cambridge); in this century, Fournier d'Albe and Charlier were the first to propose a hierarchical (now we would say fractal) model of the Universe (see Mandelbrot 1982).}. Mandelbrot (1982) developed this concept, suggesting that the galaxy distribution in the Universe is fractal, with dimension $D = 1$. In addition, in the last 20 years redshift surveys at increasing depths have revealed ever larger structures and voids (see e.g. Davis et al. 1982; de Lapparent et al. 1986, da Costa et al. 1994, Vettolani et al. 1997, and references therein). Einasto et al. (1986) found evidence that $r_0$ increases with sample volume, but they estimated that it should reach a value $\\sim 10$\\h~ for a fair sample of the Universe. Pietronero (1987) stressed that, if homogeneity is not reached, the correlation length $r_0$ cannot be taken as a measure of the clustering amplitude, and suggested a slightly but significantly different definition of the autocorrelation function. Adopting this approach, Coleman et al. (1988) reanalyzed the CfA1 redshift survey (Huchra et al. 1983), concluding that the distribution of galaxies was fractal to at least 20 \\h. On the other hand, Davis et al. (1988) found that $r_0$ increased as $r_0^{0.5}$, and not linearly with the depth as predicted for a simple fractal (see also Maurogordato et al. 1992). Others advocated the need for a multifractal approach (e.g. Balian \\& Schaeffer 1989; Martinez \\& Jones 1990; Martinez et al. 1990); this is however another issue and we will not discuss it: see for example the review by Borgani (1995) and references therein. The first redshift surveys sampled relatively small volumes, and the reality and nature of correlation amplitude variations with depth remained an open --and much debated-- question. Could these variations simply reflect fluctuations due to local structures? Were they a consequence of the fractal distribution of galaxies? Or were they an indirect consequence of the dependence of clustering on galaxy luminosity? It should be pointed out that even at a scale $\\sim 1000$ \\h~ we do not expect a {\\em perfect} homogeneity, as COBE has found evidence of anisotropy in the CMB radiation at a level $\\delta T / T \\sim 10^{-5}$ (Smoot et al. 1992), a value necessary and sufficient, at least in some standard cosmological models, to explain the formation of the observed structures. The existence of very large structures in the Universe implies that the galaxy or cluster autocorrelation function must be positive to a scale corresponding to the size of these structures; but this is not necessarily inconsistent with the measured value of the galaxy autocorrelation length $r_0 \\sim 5-6$ \\h~ for $L \\sim L_*$ galaxies (as claimed by Pietronero and collaborators), as larger structures have also a lower contrast, and the correlation of luminous matter may be significantly amplified relatively to the underlying dark matter correlation function (Kaiser 1984; Bardeen et al. 1986). Therefore, we expect that the galaxy and cluster distribution will not be perfectly homogeneous even at large scales. The claim for a fractal Universe is obviously much more stronger than that: it implies that there is no convergence to homogeneity and that it is not possible to define a mean density $\\bar{n}$ of the Universe. In the last years Pietronero and his collaborators applied statistical indicators which do not imply a universal mean density on an ever increasing number of catalogs (see e.g. Di Nella et al. 1996, Sylos Labini, Montuori \\& Pietronero 1996, Montuori et al. 1997), claiming evidence for a galaxy fractal distribution at all scales. Their results are impressive, as recently stressed by Coles (1998), but they appear to be at variance with the results of other groups who analyzed the same catalogs with standard indicators. It is clear that the situation is unsatisfactory, despite many articles and even public debates on the subject (see Pietronero et al. 1997; Davis 1997; Guzzo 1997); while the strongest support to large--scale homogeneity comes indirectly from the high level of isotropy of the CMB radiation and from two--dimensional catalogs (Peebles 1996), there is still confusion on the interpretation of the available three--dimensional data, mainly due to the difference in the statistical indicators, which do not allow a direct and quantitative comparison of the results. Therefore we analyzed the Southern Sky Redshift Survey 2 (SSRS2; da Costa et al. 1994) following the approach of Pietronero and collaborators, in order to independently check their claims and to answer to their criticisms (Sylos Labini et al. 1997, hereafter SMP) about the work of Benoist et al. (1996, hereafter Paper I). In section 2 we discuss the apparent dependence of $r_0$ on sample depth; in section 3 we describe the different statistics, and their relations; in section 4 we present and discuss our fractal analysis of the SSRS2 in comparison with the standard analysis; in section 5 we show that theoretical predictions derived from the standard CDM model can reproduce our results, and are therefore consistent with a homogeneous Universe; our conclusions are in section 6. ", "conclusions": "In this paper, we have carefully examined the claim that the Universe is a fractal, analyzing the SSRS2 and using the same statistical approach of Pietronero and collaborators. Here are our main results: \\begin{itemize} \\item the SSRS2 has a luminosity dependent correlation amplitude; \\item results obtained with the standard and fractal approach give consistent results; \\item concerning large--scale homogeneity, observational evidence is --for the moment-- consistent with theoretical predictions derived from standard CDM models of galaxy formation, and does not require a fractal Universe; \\item on the other hand, the first zero--point of the correlation function we measure in SSRS2 volume--limited subsamples increases linearly with the sample depth, and has approximately the same value of the radius of the largest sphere which can be included in the sample. This implies that the zero--point of $\\xi(s)$ is beyond 40 \\h, a lower limit set by very luminous galaxies (see Benoist et al. 1996; Cappi et al. 1998) and consistent with results from clusters (e.g. Cappi \\& Maurogordato 1992); on this point, we agree with SMP that measurements of $\\xi(s)$ beyond a scale corresponding to the radius of the largest sphere included in the sample are not reliable. \\end{itemize} We conclude that the dependence of the galaxy correlation function on the galaxy luminosity and the power still present at large scales ($\\ge 40$ \\h) can satisfactorily explain those effects interpreted by SMP as evidence of an inhomogeneous, fractal Universe." }, "9804/astro-ph9804170_arXiv.txt": { "abstract": "Observations of clusters and super clusters of galaxies have indicated that the Universe is more dominated by baryons than ever estimated in the homogeneous cosmological model for primordial nucleosynthesis. Recent detections of possibly low deuterium abundance in Lyman-$\\alpha$ clouds along the line of sight to high red-shift quasars have raised another potential difficulty that \\he4 is overproduced in any cosmological models which satisfy the low deuterium abundance constraint. We show that the inhomogeneous cosmological model with degenerate electron-neutrino can resolve these two difficulties. ", "introduction": "One of the cosmological impacts of primordial nucleosynthesis is on the universal baryon mass density $\\rho_b $. It is of significance to answer the question how much fraction of universal mass is made of ordinary matter baryons. Homogeneous Big-Bang model for primordial nucleosynthesis~\\cite{wagoner67,copi95}, assuming the standard model for light neutrino families, predicts small $\\Omega_b$, $0.03 \\le \\Omega_b h_{50}^2 \\le 0.06$, where $\\Omega_b = \\rho_b / \\rho_c$, $\\rho_c$ is the critical density which marginally closes the Universe, and $h_{50}$ is the Hubble constant $H_0$ divided by $50\\,km/s/Mpc$. However, recent observations of rich clusters and super clusters of galaxies have indicated much larger baryon fraction, 0.1 $\\le \\Omega_b h_{50}^{3/2} \\le$ 0.3~\\cite{white93,bahcall95}. Inhomogeneous Big-Bang model~\\cite{appligate85}--\\cite{mathews96}, which allows inhomogeneous baryon density distribution due to various physical processes in the early Universe, has been proposed in order to resolve this discrepancy. In this model difference in diffusion effects between neutrons and charged nuclei plays an important role in fluctuating density distribution to suppress overproduction of \\he4, and resultant $\\Omega_b$ is relaxed to $\\Omega_b \\sim 0.1$~\\cite{orito97}. However, another potential difficulty has been imposed by recent observations of deuterium absorption line in Lyman-$\\alpha$ clouds along the line of sight to high red-shift quasars~\\cite{rugers96,tytler96}. Several detections~\\cite{tytler96} among them provide too small deuterium abundance to accept concordant $\\Omega_b$ which satisfies the abundance constraints on the other light elements \\he3, \\he4 and \\li7. The observed deuterium abundance still scatters largely by one order of magnitude depending on different Lyman-$\\alpha$ systems, and there are still many error sources unclear in the analysis of abundance determination. However, if these detections are real, the abundance found there is presumed to constrain most strongly the primordial abundance because these clouds are the primitive gas which still resides in the epoch of galaxy formation and has not been processed very much in its evolutionary history. It is the purpose of this paper to propose that the inhomogeneous cosmological model with degenerate electron-neutrino can resolve these two difficulties simultaneously within the framework of the standard model for neutrino. In the next section we first discuss neutrino properties in the early Universe which can affect strongly the primordial nucleosynthesis. We then present the results of primordial nucleosynthesis calculated in both homogeneous and inhomogeneous cosmological models in sect. 3, and the $\\Omega_b$ problem and the problem of overproduction of \\he4 are discussed in details. Finally, in sect. 4, we summarize this paper. ", "conclusions": "We studied the effects of lepton asymmetry of partially degenerate electron-neutrino on the primordial nucleosynthesis. Homogeneous Big-Bang model with neutrino degeneracy parameter $\\xi_{\\nu_e}$ = 0.05 can recover the concordance between \\he4 and low deuterium abundance which was found in some Lyman-$\\alpha$ clouds along the line of sight to high red-shift quasars, but the resultant $\\Omega_b$ is less than detected in rich clusters 0.1 $\\le \\Omega_b h_{50}^{3/2} \\le$ 0.3. It was found that the inhomogeneous Big-Bang model with the same degeneracy parameter can predict $\\Omega_b$ as large as 0.22, which is in reasonable agreement with observation. This degeneracy parameter corresponds to a small chemical potential of electron-neutrino of order 10$^{-5}$ eV. It is desirable to detect the asymmetry of background neutrinos." }, "9804/astro-ph9804036_arXiv.txt": { "abstract": "We present a detailed analysis of a high resolution spectrum of the damped Ly$\\alpha$ system at $z_{\\rm abs}$~=~2.8112 toward PKS~0528-250. The absorption redshift is slightly larger than the emission redshift of the quasar. We estimate the column density of H$_2$ molecules $N$(H$_2$)~$\\sim$~6$\\times$10$^{16}$~cm$^{-2}$ and the fractional abundance of H$_2$, $f$~=~5.4$\\times$10$^{-5}$. The excitation temperature derived for different transitions suggests that the kinetic temperature of the cloud is $\\sim$200~K and the density $n$~$\\sim$~1000~cm$^{-3}$. The cloud therefore has a dimension of $\\sim$1~pc along the line of sight. Since it obscures the broad-line emission region, its transverse dimension should be larger than 10~pc.\\par We obtain upper limits on the column densities of C~{\\sc i} ($<$~10$^{12.7}$~cm$^{-2}$) and CO ($<$~10$^{13.2}$~cm$^{-2}$; $N$(CO)/$N$(H~{\\sc i})~$<$~7$\\times$10$^{-9}$). We suggest that the ratio $N$(H$_2$)/$N$(C~{\\sc i}) is a useful indicator of the physical conditions in the absorber. Simple photo-ionization models assuming solar relative abundances show that radiation fields with spectra similar to typical AGNs or starbursts are unable to reproduce all the constraints and in particular the surprisingly small $N$(C~{\\sc i})/$N$(H$_2$) and $N$(Mg~{\\sc i})/$N$(H$_2$) ratios. In view of the models we explored, the most likely ionizing spectrum is a composite of a UV-\"big bump\" possibly produced by a local starburst and a power-law spectrum from the QSO that provides the X-rays. Dust is needed to explain the production of molecules in the cloud. The amount of dust is broadly consistent with the [Cr/Zn] abundance determination. ", "introduction": "\\label{intr} QSO absorption line systems probe the baryonic matter over most of the history of the Universe (0~$<$~$z$~$\\la$~5). The so-called damped Ly$\\alpha$ (hereafter DLA) systems are characterized by a very large H~{\\sc i} column density ($N$(H~{\\sc i})~$\\ga$~2$\\times$10$^{20}$ ~cm$^{-2}$), similar to the one usually seen through local spiral disks. The case for these systems to be produced by proto-galactic disks is supported by the fact that the cosmological density of gas associated with these systems is of the same order of magnitude as the cosmological density of stars at present epochs (Wolfe 1996). Moreover the presence of heavy elements ($Z \\sim 0.1 ~ Z_\\odot$) and the redshift evolution of metallicity suggest the ongoing star formation activities in these systems (Pettini et al. 1997), while strong metal line systems have been demonstrated to be associated with galaxies at low and intermediate $z$ (e.g. Bergeron \\& Boiss\\'e 1991). It has also been shown that the profiles of the lines arising in the neutral gas show evidence for rotation (Wolfe 1996, Prochaska \\& Wolfe 1997). Whether these arguments are enough to demonstrate that DLA systems arise in large disks is a matter of debate however. Indeed simulations have shown that the progenitors of present day disks of galaxies could look like an aggregate of well separated dense clumps at high redshift. The kinematics could be explained by relative motions of the clumps with very little rotation (Haehnelt et al. 1997, Ledoux et al. 1998). Moreover, using {\\sl HST} high spatial resolution images of the field of seven quasars whose spectra contain DLA lines at intermediate redshifts (0.4~$\\la$~$z$~$\\la$~1), Le~Brun et al. (1996) show that, in all cases, at least one galaxy candidate is present within 4~arcsec from the quasar. There is no dominant morphological type in their sample: three candidates are spiral galaxies, three are compact objects and two are amorphous low surface brightness galaxies. Therefore, although the nature of the DLA systems is unclear they trace the densest regions of the Universe where star formation occurs.\\par It is thus surprising that despite intensive searches, the amount of H$_2$ molecules seems quite low in DLA systems in contrast to what is observed in our own galaxy. Two detections of H$_2$ molecules in high redshift DLA systems have been reported. Recently Ge \\& Bechtold (1997) have found strong absorptions in the $z_{\\rm abs}$~=~1.9731 DLA system toward Q~0013--004. They derive $N$(H$_2$)~=~6.9$\\times$ 10$^{19}$~cm$^{-2}$, $b$~=~15~km~s$^{-1}$, $T_{\\rm ex}$~$\\sim$~70~K and $n$(H)~$\\sim$~300~cm$^{-3}$ for a total hydrogen column density $N$(H)~=~6.4$\\times$10$^{20}$~cm$^{-2}$. This system has by far the largest H$_2$ abundance $f$~=~2$N$(H$_2$)/[2$N$(H$_2$)~+~$N$(H~{\\sc i})] $\\sim$~0.22$\\pm$0.05 observed in high z DLA systems. However the exact number should be confirmed using a higher resolution data. Other searches have led to much smaller values or upper limits ($f$~$<$~10$^{-6}$, Black et al. 1987, Chaffee et al. 1988, Levshakov et al. 1992). Levshakov \\& Varshalovich (1985) suggested that H$_2$ molecules could be present in the $z_{\\rm abs}$~=~2.8112 system toward PKS~0528--250. This claim has been confirmed by Foltz et al. (1988) using a 1~\\AA~ resolution spectrum. The latter authors derive $N$(H$_2$)~=~10$^{18}$~cm$^{-2}$, $b$~=~5~km~s$^{-1}$, $T_{\\rm ex}$~=~100~K and log~$N$(H~{\\sc i})~=~21.1$\\pm$0.3. By fitting the damped absorption together with the Ly$\\alpha$ emission from the quasar, M\\o ller \\& Warren (1993) find log~$N$(H~{\\sc i})~=~21.35. Three Ly$\\alpha$ emission-line objects have been detected within 100$h^{-1}$~kpc from the quasar by M\\o ller \\& Warren (1993) and confirmed by Warren \\& M\\o ller (1996) to have redshifts within 200~km~s$^{-1}$ from the redshift of the DLA system ($z_{\\rm abs}$~=~2.8112 as measured on the Ni~{\\sc ii} lines by Meyer \\& York 1987). The widths of the Ly$\\alpha$ emission lines are very large ($>$~600~km~s$^{-1}$) and continuum emission could be present (Warren \\& M\\o ller 1996); this suggests that the gas is not predominantly ionized by the quasar and that star-formation may occur in the clouds, a conclusion reached as well by Ge et al. (1997). The proximity of the quasar makes the case difficult however and careful analysis is needed.\\par In this paper we use much higher spectral resolution data to reanalyze the molecular lines in this system. We present the observations in Section~2, the results in Section~3 and discuss the low $N$(C~{\\sc i})/$N$(H$_2$) ratio inferred in Section~4. ", "conclusions": "\\label{s4} By fitting the different H$_2$ transitions that we detect in a high resolution spectrum of PKS~0528--250, we derive log~$N$(H$_2$)~$\\sim$~16.78 and $T_{\\rm ex}$~$\\sim$~200~K that is most certainly the true kinetic temperature of the gas. For this temperature, the relative populations of rotational levels 0 to 4 indicate that the density is of the order of 1000~cm$^{-3}$. Therefore the dimension of the neutral cloud along the line of sight is less than 1~pc. It must be noticed that the damped absorber must cover the broad line region. Indeed there is no residual flux in the bottom of the Ly$\\alpha$ absorption line which completely absorbs the Ly$\\alpha$ emission from the quasar over more than 5000~km~s$^{-1}$ (M\\o ller \\& Warren 1993). Since the dimension of the BLR in QSOs can be approximated as $R$~$\\sim$~0.3~$L_{46}^{0.5}$ where $L_{46}$ is the bolometric luminosity in units of 10$^{46}$~erg~s$^{-1}$ (Collin, private communication). The radius of the BLR in PKS~0528--250 is thus of the order of 10~pc. The transverse dimension of the damped cloud must be thus larger than 10~pc and the cloud must be quite flat. We can derive an upper limit on the transverse dimension of the cloud by interpreting the non-dection of redshifted 21 cm absorption (Carilli et al. 1996) as an effect of partial covering factor of the continuum radio emission by the cloud. Indeed the size of the radio source is 1~arcsec or 5$h^{-1}_{75}$~kpc. If the spin temperature is equal to the kinetic temperature we derived in sect. 3.2, this implies that the covering factor should be less than 0.3 and thus the radius of the cloud along the transverse direction is less than 1~kpc. We determine upper limits for the C~{\\sc i}, Mg~{\\sc i} and CO column densities (log~$N$(C~{\\sc i})~$<$~12.7, log~$N$(Mg~{\\sc i})~$<$~12.8 and log~$N$(CO)~$<$~13.2). Based on simple photoionization models we conclude that (i) no simple model can reproduce at the same time the low $N$(C~{\\sc i})/$N$(H~{\\sc i}), $N$(Mg~{\\sc i})/$N$(H~{\\sc i}) ratios and the presence of molecules at the level observed; (ii) steep spectra are required to reproduce the low $N$(C~{\\sc i})/$N$(H~{\\sc i}) and $N$(Mg~{\\sc i})/$N$(H~{\\sc i}) ratios but they predict temperatures smaller than 100~K in conflict with the excitation temperature derived from the H$_2$ transitions; (iii) the only way to keep the temperature as high as 200~K is to allow for some X-ray flux heating the gas; (iv) in the framework of these models, dust is needed to produce the observed amount of molecules; (v) the gas phase abundances of carbon and especially magnesium should be smaller than 0.1 of solar. In view of the models we explored, the most likely ionizing spectrum is a composite of a UV-\"big bump\" possibly produced by a local starburst and a power-law spectrum from the QSO that provides the X-rays." }, "9804/astro-ph9804292_arXiv.txt": { "abstract": "The effects that large scale fluctuations had on small scale isothermal modes at the epoch of recombination are analysed. We find that: (a)~Albeit the fact that primordial fluctuations were at this epoch still well in the linear regime, a significant nonlinear radiation hydrodynamic interaction could have taken place. (b)~Short wavelength isothermal fluctuations are unstable. Their growth rate is exponential in the amplitude of the large scale fluctuations and is therefore very sensitive to the initial conditions. (c)~The observed CMBR fluctuations are of order the limit above which the effect should be significant. Thus, according to their exact value, the effect may be negligible or lead to structure formation out of isothermal fluctuations within the period of recombination. (d)~If the cosmological parameters are within the prescribed regime, the effect should be detectable through induced deviations in the Planck spectrum. (e)~The sensitivity of the effect to the initial conditions provides a tool to set limits on various cosmological parameters with emphasis on the type and amplitude of the primordial fluctuation spectrum. (f)~Under proper conditions, the effect may be responsible for the formation of sub-globular cluster sized objects at particularly high red shifts. (g)~Under certain circumstances, it can also affect horizon sized large scale structure. ", "introduction": "The problem of structure formation in the Universe has probably been one of the foremost and most studied question in cosmology. Perhaps the greatest achievement of cosmology was the prediction and following discovery of the Cosmic Microwave Background Radiation -- the CMBR (Penzias \\& Wilson, 1965) and its anisotropy (Smoot et al. 1992). Not only did it provide support for the Big Bang theory, it proved that structure is a result of small fluctuations growing into large inhomogeneities and not vice versa. The fact that the observed fluctuations in the CMBR are small ($\\Delta T/T \\sim 10^{-5}$), naively implies that one can treat the fluctuation within the linear approximation when modeling the evolution of structure before the time of last photon scattering, i.e., one can assume that fluctuation modes of different wavelengths are completely decoupled from each other before radiation-matter decoupling. The first to develop a linear theory for the perturbed Fridmann - Robertson - Walker metric was Lifshitz (1946), and it can be found in various text books such as Weinberg (1972), Peebles (1980) and Kolb \\& Turner (1990). The theory has since then been applied to describe cosmologies with various components and various initial parameters. Baryonic matter, radiation and other massless particles, cold (massive) or hot (light) dark matter are some of the ingredients that enter the primordial soup. While other parameters such as the Hubble constant $H_0$, the vacuum energy through the cosmological constant $\\Lambda$, and the initial spectrum affect the evolution. A review of various current cosmological models, their evolution and implication can for example be found in White et al. (1994) with emphasis on the microwave background radiation and in Primack (1997). Less current reviews that cover the basic principles of structure formation are found in the aforementioned textbooks. One of the major parameters that affect the qualitative behaviour of our universe and is highly relevant to the present paper, is the type and form of the primordial spectrum of fluctuations. The fluctuations are generally classified into curvature (or adiabatic) and isocurvature (or isothermal) fluctuations. The first arise naturally in inflationary scenarios (e.g., Liddle and Lyth, 1993, and ref. therein). They are fluctuations of both the baryonic fluid and the radiation and they propagate at the adiabatic speed of sound that is equal to $c/\\sqrt{3}$ if the radiation energy density dominates. These waves are found to decay on scales smaller than the Silk scale, namely, scales comparable to galaxy sized objects (Silk, 1967). Therefore, a top-down structure formation is a natural consequence of adiabatic perturbations. Small scale objects can form after recombination (and initiate a bottom-up scenario) only if one adds the undamped perturbation of a cold component as is the case in the standard CDM model (Peebles 1982, Blumenthal et al. 1984 and Davis et al. 1985) and its variants. Isocurvature (or isothermal) fluctuations on the other hand, are a natural consequence of topological defects formed in phase transitions (e.g, strings, monopoles and textures) or if more than one field contributes significantly to the energy density during inflation. They correspond to fluctuations that alter the entropy density but not the energy density. Unlike the first type of fluctuations, these do not suffer from Silk damping. Consequently, objects with a mass as small the post-recombination Jeans scale, namely the size of globular clusters, can form after decoupling. It is interesting and important to note that even if the primordial spectrum was purely adiabatic, isothermal perturbations of a given wavelength are formed as the second order effect of Purcell clustering from adiabatic waves of similar wavelengths (Press \\& Vishniac, 1979). Through this effect, noninteracting particles (baryons) that are viscously coupled to a stochastically oscillating background gas (radiation fluid) undergo secular clustering. Originally, it was thought that the effect can produce a bottom-up scenario even from a pure baryon model with an adiabatic spectrum. However, the typical post-recombination Jeans scale amplitude of the isothermal waves will only be $\\rho_{iso}/\\rho\\sim (10^{-2} - 1) \\times \\left( \\delta \\rho_{ad} / \\rho\\right)^2$, with $\\delta \\rho_{ad} /\\rho$ the typical adiabatic fluctuations at horizon crossing. Namely, if the primordial spectrum is flat and adiabatic, the typical isothermal fluctuation at recombination is roughly $\\delta \\rho_{iso} /\\rho \\sim 10^{-10}-10^{-8}$. For a flat isothermal primordial spectrum, one should expect typical amplitudes of $\\rho_{iso}/\\rho \\sim 10^{-5} - 10^{-4}$. The smallness of nonlinear effects such as the Purcell clustering (or shock waves which are of an even higher order) led to the consensus that the evolution of the fluctuations can be treated linearly and modes of different wavelengths are decoupled from each other. This delays the nonlinear treatment to the time when radiation does not play any dynamic role anymore. At face value, it certainly appears to be the case as $\\delta\\rho /\\rho$, $v/c$, and $\\delta T/T$ are all much smaller than unity. One should nevertheless be extremely careful when assuming linearity, especially in view of the fact that not all of the dimensionless parameters are actually smaller than unity. One of the dimensionless numbers that appears in the solution of the nonlinear fluctuations' equations of motion and that is not small at all is the root of the radiation to gas pressure ratio. Just before recombination one finds that $\\sqrt{P_{rad}/P_{gas}}\\approx{10}^{5}$! Consequently, we expect that the interaction between the radiation and matter at this period will have profound impact on the evolution of the fluctuations. In this paper we examine the linear hypothesis and its validity by adding the force large scale perturbations exert on short wavelength waves. We begin in \\S2 by overviewing the problem of solving the nonlinear equations of motion and estimating the effect with a very simple analysis. We proceed in \\S3 to write the Newtonian equations describing the evolution of short wavelength isothermal modes. In \\S4 we analyse the simplified solution, while in \\S5 we proceed to estimate the effect in a few cosmological scenarios. In \\S6 \\& \\S7, we study the possible ramifications to structure formation and study the possibility of measuring and using this effect in the study of cosmological parameters. In \\S8 we show that the effect can influence large scale structure as well. ", "conclusions": "We have analysed the nonlinear growth of small scale isothermal fluctuation through the induced effect that large-scale perturbations have. The equations are solved through the separation of the large scale inhomogeneities into smaller regions where they can be considered uniform, and the shear $\\Delta v_0$ exerted by the large scales is constant in space. This can be performed because the waves that interest us are waves with very small wavelengths and a very small propagation speed, such that over the integration time, the waves practically remain in the same region of space. Small scale isothermal waves are found to be unstable if the large scale shear velocity between the ``photon fluid'' and the baryon fluid is larger than roughly twice the isothermal speed of sound at the time of recombination. A period when the logarithmic derivative of the opacity with respect to the density does not vanish but is $-1/2$ instead. When the shear is smaller than the critical value, the solution found is qualitatively the same as the solution for the completely decoupled equations. The shear in this case may induce some quantitative corrections but the behaviour of the system does not change. The behaviour of the system is however radically different if the shear is larger than a critical value. For wavelength shorter than the roughly the Jeans scale, the growth rate increases with $k$ and it saturates when $\\tau_{e\\gamma}=\\tau_o$, i.e., at a scale which is a few times smaller than the Jeans scale. The maximal growth rate attained is: \\begin{equation} r_{\\infty} = {\\alpha \\over 2} {\\Delta v_0 \\over v_s} {1\\over \\tau_e{\\gamma}}. \\end{equation} If $r_{\\infty}$ is much larger than all other rates in the system, it can be easily integrated over the duration of decoupling $\\Delta t$. The growth factor obtained is then: \\begin{equation} {\\delta_1\\over\\delta_0} \\approx \\exp (G_{\\infty}) = \\exp \\left(\\left\\langle {\\alpha \\over 2} {\\Delta v_0 \\over v_s} {1\\over \\tau_e{\\gamma}}\\right\\rangle_t \\Delta t\\right), \\end{equation} with the brackets denoting a temporal average. The isothermal waves are unstable because their sound speed is very small. Consequently, the acoustic energy in a wave with a given amplitude is very small and the amount of work needed to increase its amplitude isn't large, in fact, it is so small that nonlinear effects start to take place already at $\\delta\\rho/\\rho\\sim 10^{-5}~{\\rm to}~10^{-4}$! Even more surprising is the fact the the typical shears found in typical cosmological scenarios is {\\em exactly} at the verge of having nonlinear effects take place! If the amplitude of the large scale fluctuations would have been an order of magnitude larger, then any isothermal perturbations, however small, could have been amplified right into the nonlinear regime. On the other hand, if the typical amplitude would have been a few times smaller, the effect would have been completely meaningless. For typical cosmological scenarios normalized to COBE, the growth parameter is roughly $G_{rms} \\sim 1 - 5$, e.g., for an $\\Omega=1,~\\Omega_b=0.05,~h=.66$ CDM model one finds $G_{rms} \\approx 5$, if the spectral index is tilted to $n=0.9$, one finds $G_{rms}\\approx 4$, if a hot component is added at a 20\\% level to the untilted model, it falls to $G_{rms} \\sim 1$. The fate of the regions that do reach nonlinearity is still an open question. Do these regions collapse and form black holes? Do they form an early generation of massive stars? Do they release enough energy to the environment and affect it, or perhaps, even form structure on much larger scales? Although some speculations of what might occur do exist in the context of early small scale structure formation, it is not entirely clear. Moreover, nonlinear hydrodynamic simulations is probably unavoidable if we are to really solve the problem. The only relatively certain consequence is that the dissipation taking place after recombination heats the matter and it is likely to ionise it, raise its temperature to the ionisation temperature and leave an imprint on the CMBR in the form of a deviation from a Planckian spectrum. In scenarios where $G_{rms}\\gtrsim 5$, isothermal perturbations with an amplitude of as low as $\\delta\\rho_{iso}/\\rho\\sim10^{-6}$ will be in fact theoretically detectable in the future. An amplitude of $\\delta\\rho_{iso}/\\rho\\sim10^{-4}$ will leave 1\\% of the background sky with a detectable deviation. This fraction can change considerably if the density perturbations are not Gaussian. The detection of deviations from a Planckian spectrum or the placing of an upper limit for them can result with interesting implications. Any deviations found will first of all directly prove the existence of primordial isothermal fluctuations. Second, such a detection will place stringent limits on cosmological parameters, as only those scenarios that produce a large enough $G_{rms}$ are capable of producing nonlinear structure, and because the amount of nonlinearity is extremely sensitive to $G_{rms}$, it is also sensitive to the exact cosmological parameters. Third, if the cosmological parameters are known with an large accuracy (for example, through the fitting of the CMBR spectrum) then the primordial isothermal perturbation spectrum at very large $k$'s can be estimated. Even if no detection of a deviation from a Planckian spectrum is found, one can place limits on cosmological parameters. Moreover, if in the future it will be found that the cosmological parameters actually correspond to a high $G_{rms}$ model, one will be able through the lack of detection of a $y$-parameter to place extremely powerful limits on the amplitude of the isothermal component of the primordial spectrum. Through the Press \\& Vishniac effect, limits can also be placed on the adiabatic spectrum. Another interesting implication is the possibility of transferring enough energy from the large to the small scales and considerably change the amplitude of the large-scales. This possibility relies on whether a significant volume fraction can be amplified out of the linear regime and on the maximum dissipation rate of nonlinear isothermal waves. Under favourable circumstances, the large scale amplitude can be significantly reduced such that $G_{rms}$ calculated from the observed large scale fluctuations would only be of order unity. The fact that the observed fluctuations in the CMBR correspond to values of this order raises a very interesting question. Why do the values of $G_{rms}$ corresponding to the observed fluctuations happen to fall in a small region around unity? Is it because the universe was created with a fluctuation spectrum corresponding to this region, or, is it because $G_{rms}$ that evolved from the initial cosmological parameters was actually larger but it was naturally reduced to the oberved value? One can summarise the several plausible scenarios according to both the value of $G_{rms}$ predicted by the cosmological model and the amplitude $\\delta = \\delta \\rho_{iso}/\\rho \\approx 10^{-10} - 10^{-5} $ of the sub Jeans scale isothermal fluctuations before the advent of recombination, these possibilities are: \\begin{enumerate} \\item If $G_{rms}$ is of order unity or less, the typical growth of isothermal waves is at most a few $e$-folds. The effect will be insignificant as the predicted micro degree size fluctuations are neither fluctuations of the temperature nor on a scale measurable in the near future. Moreover, there are no implications at all to structure formation. \\item For a value of $G_{rms} \\gtrsim - (\\ln \\delta)/3 \\approx 4 - 8$, the effect will be measurable as patches in the CMBR with a distorted Planckian spectrum. As long as $G_{rms} \\lesssim - \\ln \\delta \\approx 10-20$, only a small fraction of the universe would have reached nonlinearity and formed small scale structure by the end of recombination. Under certain circumstances however, it can lead to large scale structure formation through explosive amplification. \\item For a value of $G_{rms}$ such that $ G_{rms} \\gtrsim - \\ln \\delta \\approx 10-20$, a large fraction of the universe reaches nonlinearity by the end of recombination and small scale structure is subsequently formed in most of the universe soon after recombination. The large scale spectrum is damped by transferring a significant amount of energy to the small scales, thus, the $G_{rms}$ measured from the damped spectrum is smaller than the $G_{rms}$ calculated when neglecting this process. In some cases, $G_{rms}$ will be reduced to a values of order unity and it will mimic values on the boundary between the first and second scenarios. Note that it will not affect fluctuations in components that decouple from the matter-radiation fluid before recombination (e.g. dark matter fluctuation). \\end{enumerate} The analysis presented here is the first step in the investigation of the radiation-matter interaction instability. More accurate integration is needed to improve the actual transfer function for isothermal waves. The analysis here was restricted to the evaluation of $G_{rms}$ and it does not include the actual rates which are $k$ dependent. This approximation overestimates the rate of growth of finite sized $k$'s. In the present paper we have simulated the freezing of the instability due the freezing of recombination by stopping the integration abruptly at a given $z$. However, the equilibrium conditions and therefore the switching off of the effect depend on the isothermal wavelength as well. By assuming this assumption, we have actually underestimated the contribution from waves of order the Jeans size. Many of the possible implications depend on the quantitative behaviour of of the nonlinearities once they are reached. A numerical hydrodynamic study of large amplitude waves will certainly help us understand of the fate of the nonlinear objects and the possible implications they have on large scale structure as well." }, "9804/astro-ph9804017_arXiv.txt": { "abstract": "I summarize some recent models and ideas for the formation of axisymmetrical structures of planetary nebulae and the three rings of SN 1987A, as follows. (a) I review the general role of binary companions, including brown dwarfs and planets. (b) I propose a mechanism for axisymmetrical mass loss on the AGB that may account for the axially symmetric structures of elliptical planetary nebulae and that operates for slowly rotating AGB stars, $10^{-4} \\Omega _{\\rm Kep} \\lae \\Omega \\lae 10^{-2} \\Omega_{\\rm Kep}$, where $\\Omega_{\\rm Kep}$ is the equatorial Keplerian angular velocity. (c) I propose a model for the formation of the two outer rings of SN 1987A, which is based on the numerical simulation of Soker (1989), and discuss a mechanism for their displacement from the exploding star. ", "introduction": "Scanning through recent images of SN 1987A (e.g. Burrows {\\it et al.} 1995) and through catalogs of planetary nebulae (PNs; e.g., Acker {\\it et al.} 1992; Schwarz, Corradi, \\& Melnick 1992; Manchado {\\it et al.} 1996) we find that the circumstellar media of many stars at their final nuclear burning phase have axisymmetrical, rather than spherical, structures. Axisymmetrical PNs which have two lobes with a morphological ``waist'' between them are termed ``bipolar PNs'' (also ``butterfly'' or ``bilobal''), while PNs which have a more elliptical than bilobal structure are termed elliptical PNs (Schwarz, Corradi, \\& Stanghellini 1993). The axisymmetrical structures of most PNs led to a debate on whether elliptical PNs can be formed through single-stellar evolution, or whether a binary companion is necessary (e.g., Fabian \\& Hansen 1979; Livio 1982, 1998; Livio, Salzman, \\& Shaviv 1979; Webbink 1979; Morris 1981; Zuckerman \\& Gatley 1988; Pascoli 1992; Iben \\& Livio 1993; Soker 1997, 1998a; Balick {\\it et al.} 1994; Pottasch 1995; % Pollacco \\& Bell 1997; Corradi {\\it et al.} 1996; Kastner {\\it et al.} 1996). In the last decade this debate was extended to the formation of the nonspherical explosion and three rings of SN 1987A. In many PNs, as well as in the three rings of SN 1987A, there are displacements of the nebulae from the central stars, which hint at the interaction of the progenitors with wide binary companions, with close binaries having eccentric orbits, or with the ISM ($\\S 5$). In a recent paper (Soker 1997) I suggest that four main evolutionary routes determine the degree of asymmetry of the axially symmetric structures of PNs. I then classify 458 PNs according to the process which caused their progenitors to blow axisymmetrical winds. The classification is based primarily on the morphologies of the different PNs, assuming that binary companions, stellar or substellar, are necessary for axisymmetrical mass loss on the AGB. The four evolutionary classes, according to the binary-model hypothesis, are: \\newline (a) Progenitors of planetary nebula which did not interact with any companion, and therefore they rotate extremely slowly when reaching the AGB. These amount to $\\sim 10 \\%$ of all planetary nebulae. \\newline (b) Progenitors which interact with stellar companions which avoided a common envelope, $11 ^{+ 2}_{-3} \\%$ of all nebulae. These form bipolar PNs, as is the case in symbiotic nebulae (Morris 1990; Schwarz \\& Corradi 1992; Soker 1998a). % \\newline (c) Progenitors which interact with stellar companions via a common envelope phase, $23^{+11}_{-5} \\%$ of all nebulae. These form extremely asymmetrical structures, i.e., tori, elongated elliptical PNs, and in some cases bipolar PNs. \\newline (d) Progenitors which interact with {\\it substellar} (i.e., planets and brown dwarfs) companions via a common envelope phase, $56^{+5}_{-8} \\%$ of all nebulae. These form elliptical PNs with relatively small deviation from sphericity. \\newline These numbers are compatible with other studies (e.g., Yungelson, Tutukov, \\& Livio 1993; Han, Podsiadlowski, \\& Eggleton 1995). In $\\S 2$ I discuss the problem of angular momentum of AGB stars, which suggests that to account for the $\\sim 60 \\%$ elliptical PNs either there are many planetary systems ($\\S 3$; Soker 1996; 1997) or there is a mechanism to induce axisymmetrical mass loss from very slowly rotating AGB stars. Such a model for singly evolved very slowly rotating AGB stars is the mechanism of mode-switch to nonradial oscillations, proposed by Soker \\& Harpaz (1992). In $\\S 4$ I propose yet another model (Soker 1998c) which may operate in singly evolved AGB stars. This model is based on both magnetic activity and radiation pressure on dust. In $\\S 5$ I propose a model for the two outer rings of SN 1987A. ", "conclusions": "" }, "9804/astro-ph9804221_arXiv.txt": { "abstract": "We investigate a one-zone chemophotometric evolution model of disk-disk galaxy mergers in order to clarify whether or not galaxy mergers with the widely spread merging epoch can reproduce reasonably well the observed small scatter of the color-magnitude ($C-M$) relation in cluster ellipticals at low and intermediate redshift ($z<1$). We consider that merger progenitor disks begin to consume interstellar gas at moderate rate from $z \\sim 5$ and then merge to form an elliptical with the secondary starburst at $z=z_{\\rm merge}$. We find that even if the epoch of galaxy merging is rather extended ($0.3 2.0$ (Arag\\'on-Salamanca et al. 1993; Ellis et al. 1997; Stanford, Eisenhardt, \\& Dickinson 1998). This classical picture of coeval elliptical galaxy formation also appears to be supported by small redshift evolution of both the mass-to-light-ratio (van Dokkum \\& Franx 1996) and the $\\rm Mg_{2} - \\sigma$ relation (Ziegler \\& Bender 1997). The considerably tight $C-M$ relation (Bower, Lucey, \\& Ellis 1992) and the Fundamental Plane (e.g., Djorgovski \\& Davis 1987) at the present epoch, and redshift evolution of the slope and the zero-point of the $C-M$ relation (Kodama \\& Arimoto 1997; Gladders et al. 1998; Kodama et al. 1998) furthermore seem to support the coevality of elliptical galaxy formation. An increasing number of recent observational results, however, shed a strong doubt on this long-standing view of elliptical galaxy formation, and suggest that there is great variety of star formation history among elliptical galaxies, such as the epoch of major star formation, the duration and efficiency of star formation (Worthey, Faber, \\& Gonzalez 1992; Faber et al. 1995; Worthey, Trager, \\& Faber 1996). In particular, Faber et al. (1995) suggested that the `apparent age spread', which is inferred from the combination of line index analysis of elliptical galaxies, amounts to $\\sim$ 10 Gyr. Schweizer \\& Seitzer (1992) found that in merger remnants with morphologically fine structures, the last merging epoch, which corresponds to elliptical galaxies formation, ranges from 4.6 Gyr to 8.0 Gyr ago. These observed spread in `apparent mean age' seem to be inconsistent with the aforementioned coevality of elliptical galaxy formation expected mainly from the redshift evolution of the $C-M$ relation. The purpose of this paper is to give a plausible answer to the above apparent inconsistency in the epoch of elliptical galaxy formation. We adopt the merger scenario of elliptical galaxy formation (e.g., Toomre \\& Toomre 1972) and thereby investigate to what degree the difference in the epoch of major galaxy merging (i.e., the epoch of elliptical galaxy formation) can be allowed to preserve the observed small scatter of the $C-M$ relation of cluster ellipticals ($\\sim 0.05$ mag) at $z$ = 0 (Bower et al. 1992), 0.55 (Ellis et al. 1997), and 0.895 (Stanford et al. 1998). We find that owing to the age-metallicity conspiracy proposed by Worthey et al. (1996), the observed small scatter in the $C-M$ relation can be reproduced reasonably well even in star-forming galaxy mergers with the widely spread merging epoch. This result accordingly reinforces the recent results of Kauffmann \\& Charlot (1998), in which the tight $C-M$ relation can be successfully reproduced by merger scenario of elliptical galaxy formation based on a hierarchical clustering scenario. This result furthermore implies that the previously suggested interpretation of the tightness of the $C-M$ relation at low and intermediate redshift ($z < 1$) is {\\it not unique}, thus that the formation epoch of elliptical galaxies can be more widely spread than the classical passive evolution picture predicts. Thus, the above apparent inconsistency in the interpretation of the $C-M$ relation can be due primarily to the fact that previous studies claiming the coevality of elliptical galaxy formation did not explore so extensively possible variety in star formation history of elliptical galaxies. ", "conclusions": "The present study predicts that even if the epoch of major galaxy merging (i.e., the epoch of elliptical galaxy formation) is rather spread, both the tightness and the slope of the $C-M$ relation can be kept owing to the age-metallicity conspiracy originally proposed by Worthey et al (1996). This result accordingly provides a heuristic explanation for the result of Kauffmann \\& Charlot (1998) in which the tight $C-M$ relation has been already reproduced in the merger scenario of elliptical galaxy formation based on the hierarchical clustering model. The conclusions derived in the present study however {\\it seem} to be inconsistent with those derived in previous ones on the redshift evolution of the slope, zero-point and tightness of the $C-M$ relation of elliptical galaxies (Bower et al. 1992; Arag\\'on-Salamanca et al. 1993; Kodama \\& Arimoto 1997). In particular, the present numerical results {\\it seem} to disagree with those of Kodama \\& Arimoto (1997) and Kodama et al. (1998) (see also Gladders et al. 1998), which claim that the considerably less significant evolution of the slope of the $C-M$ relation rejects the age spread larger than 1 Gyr among elliptical galaxies. The apparent disagreement between the present study and the previous ones (e.g., Kodama \\& Arimoto 1997; Kodama et al. 1998) is due essentially to the fact that the previous studies inevitably have over-interpreted the redshift evolution of the $C-M$ relation owing to the ad hoc assumption adopted in the previous studies. Although the previous studies are considerably sensible and valuable, it is important to point out the ad hoc assumptions adopted in the previous studies and thereby clarify the reason why the present conclusions are not consistent with those derived by the previous studies of Kodama \\& Arimoto (1997) and Kodama et al. (1998). The following three are the ad hoc assumptions which inevitably lead the previous studies to draw the strong and general conclusion that formation of elliptical galaxies (especially in the cores of clusters) are {\\it as a whole} coeval and occurred at high redshift. First is that elliptical galaxies are formed by {\\it only one} initial starburst. Owing to this assumption, time evolution of global colors of elliptical galaxies depends exclusively on the epoch of initial burst of star formation (i.e., the epoch of elliptical galaxy formation in the previous study). As a result of this, the age difference between elliptical galaxies (i.e., the difference of the epoch of elliptical galaxy formation in the previous studies) can be more clearly reflected on the redshift evolution of the slope of the $C-M$ relation in the previous studies. Accordingly the observed less significant evolution of the slope of the $C-M$ relation is more likely to be interpreted as an evidence that supports the coevality of elliptical galaxy formation. It is certainly reasonable to claim that the observed evolution of the $C-M$ relation reject the `pure age' sequence model which demands that less luminous ellipticals have younger age. However, it seems not to be so reasonable to draw strong and general conclusion that elliptical galaxies are formed at $z>2$ {\\it only} from the redshift evolution of the $C-M$ relation. Considering the first ad hoc assumption in the previous studies, what is more accurate and plausible interpretation on the observed evolution of the $C-M$ relation is just that the formation of {\\it stellar populations} in {\\it some} elliptical galaxies in the cores of {\\it some} clusters (not the formation of galaxies with structural and morphological properties similar to those of ellipticals) can be coeval and occurred at higher redshift ($z > 2$). The second is that an elliptical galaxy in a cluster of galaxies at higher redshift is a precursor of an elliptical galaxy in a cluster at lower redshift. The third is that a cluster of galaxies observed at higher redshift is a precursor of a cluster of galaxies at lower one. These two ad hoc assumptions actually enable us to discuss the origin of elliptical galaxies in a more general way and thus lead us to draw more general conclusions on the formation epoch of elliptical galaxies. However, since there are no observational evidences which can provide the firm physical basis for the above assumptions at least now, it is questionable to give any general conclusions on the coevality of elliptical galaxy formation. Thus, these three assumptions adopted in the previous studies inevitably lead them to provide the strong and general conclusion that formation of elliptical galaxies are coeval and occurred at higher redshift. The present study, on the other hand, does not adopt the above three ad hoc assumptions, and rather relaxes these assumptions. Furthermore the present study instead allows both continuous and moderate star formation (not strong initial starburst) and the secondary starburst associated with galaxy merging, and assumes that the epoch of morphological transformation (into ellipticals) does not necessarily coincide with the epoch of galaxy formation (i.e., the epoch when the star formation begins). The evolution of the $C-M$ relation in the present study consequently does not depend so strongly on the difference in the formation epoch between elliptical galaxies (i.e., the epoch of major galaxy merging with the secondary starburst). As a result of this, the present merger model predicts that even if the formation epoch of elliptical galaxies (i.e., the epoch of galaxy merging) are rather spread, both the slope and tightness of the $C-M$ relation can be kept. Thus, the essential reason for the aforementioned apparent disagreement on the coevality of elliptical galaxy formation is that the present study does not adopt the above three ad hoc assumptions whereas the previous studies do. The interpretation on the redshift evolution of the $C-M$ relation in each model can depend strongly on the assumptions adopted by each model. It is safe for us to say that it is not clear, at least now, which of the two different conclusions on the coevality of elliptical galaxy formation is more plausible and reasonable. However, considering the above three ad hoc assumptions adopted in the previous studies, what is more reasonable interpretation on the redshift evolution of the $C-M$ slope is that only {\\it stellar populations} (not elliptical morphology) in {\\it some} ellipticals located in the cores of {\\it some} clusters of galaxies are formed at higher redshift. We should not draw any {\\it general} conclusions from the redshift evolution of the slope of the $C-M$ relation. Environmental difference of stellar populations (in particular, the existence of intermediate-age population) in early-type galaxies has been already indicated by a number of observational studies (e.g., Bower et al. 1990; Rose et al. 1994; Mobasher \\& James 1996). On the other hand, the tightness and the slope of the $C-M$ relation of early-type galaxies are observationally revealed not to depend so strongly on galaxy environments. These two apparently inconsistent observational results on spectrophotometric properties of elliptical galaxies have called into the following question: ``Why does not the $C-M$ relation of early-type galaxies depend strongly on galaxy environments (e.g., between rich clusters and poor ones), though stellar populations and star formation histories in early-type galaxies probably depend on galaxy environments?'' To give a plausible answer for this question seems to be important because the above apparently inconsistent observational results give us valuable information both on the environmental difference in the details of physical processes of elliptical galaxy formation and on a certain mechanism for the tight $C-M$ relation. However, no extensive theoretical studies have yet addressed the above important question. The present study has shown that the age-metallicity conspiracy, which is achieved by younger and more metal-enriched stellar populations created in the secondary starburst of galaxy mergers, allows both the apparent age spread of elliptical galaxies and the tightness of the $C-M$ relation. This result seems to provide a clue to the above question. Since the {\\it real} question concerning the tight $C-M$ relation is not to determine the typical epoch of elliptical galaxy formation but to give a convincing explanation for the reason why possible diversity in star formation histories of elliptical galaxies can allow the tight $C-M$ relation, more extensive theoretical studies including more variety of star formation history of elliptical galaxies and its likely dependence on galaxy environments are certainly worth for our deeper understanding of the origin of the tight $C-M$ relation. The present numerical results are consistent with recent observational results which suggest that coeval elliptical galaxy formation with initial starburst at higher redshift ($z > 2.0$) is not promising. Kauffmann, Charlot, \\& White (1996) revealed that only about one-third of bright E/S0 galaxies in the sample of Canada-France Redshift Survey were already in the passive evolution phase at $z \\sim 1.0$. Franceschini et al. (1997) found a remarkable absence of early-type galaxies at $z > 1.3$ in the $K$-band selected sample of early-type galaxies in the Hubble Deep Field (HDF), which suggests either that early-type galaxies are formed by galaxy merging with less prominent star formation or that a dust-polluted interstellar gas obscures forming elliptical galaxies till $z = 1.3$. Zepf (1997) demonstrated that strong deficit of galaxies with extremely red colors in the HDF means that the formation epoch of typical elliptical galaxies is $z < 5.0$. Sample galaxies in these studies are selected from {\\it field ellipticals}, which possibly have star formation histories different from those of {\\it cluster ellipticals}. Accordingly, it might not be plausible to derive strong conclusions on the formation epoch of ellipticals. However these observational results together with the present results seem to support the merger scenario which can naturally predict that the epoch of elliptical galaxy formation is rather extended ranging from high redshift to moderate one. Thus we have succeeded in pointing out that even if the epoch of elliptical galaxy formation (i.e., the epoch of major disk-disk galaxy merging, in this study) is rather widely spread, the tightness of the $C-M$ relation at low and intermediate redshift can be kept reasonably well. This result suggests that coevality of elliptical galaxy formation, which has been conventionally believed in the classical passive evolution picture, is {\\it not unique} interpretation for the small scatter of the $C-M$ relation. This furthermore implies that {\\it only} the tightness of the $C-M$ relation at a given redshift {\\it does not necessarily} give strong constraints on the formation epoch of elliptical galaxies. Worthey et al. (1996) have already pointed out that the age-metallicity conspiracy can keep both the tightness of the Fundamental Plane and that of the $C-M$ relation in elliptical galaxies. The present numerical study, which is different from the Worthey's single stellar population analysis, has confirmed that the proposed age-metallicity conspiracy can actually operate to keep convincingly the tightness of the $C-M$ relation of ellipticals formed by disk-disk galaxy mergers. The present chemophotometric evolution model is, however, not so elaborated and realistic in that this model neither includes continuous gas accretion/merging expected from a specific cosmology (e.g., Baugh, Cole, \\& Frenk 1996; Kauffmann \\& Charlot 1998) nor considers important dynamical effects of galaxy merging on chemical and photometric evolution of galaxies (Bekki \\& Shioya 1998). Accordingly it is our future study to confirm that the results derived in the present preliminary study can hold even for more sophisticated and realistic merger models. Furthermore, we should check whether or not observed redshift evolution of other fundamental relations such as the $\\rm Mg_{2} - \\sigma$ relation (Ziegler \\& Bender 1997), the Fundamental Plane (van Dokkum \\& Franx 1996), and the abundance ratio of [Mg/Fe] can be also reproduced self-consistently by our future merger model." }, "9804/hep-ph9804378_arXiv.txt": { "abstract": "In this paper we present the first analytic model for vorton formation. We start by deriving the microscopic string equations of motion in Witten's superconducting model, and show that in the relevant chiral limit these coincide with the ones obtained from the supersonic elastic models of Carter and Peter. We then numerically study a number of solutions of these equations of motion and thereby suggest criteria for deciding whether a given superconducting loop configuration can form a vorton. Finally, using a recently developed model for the evolution of currents in superconducting strings we conjecture, by comparison with these criteria, that string networks formed at the GUT phase transition should produce no vortons. On the other hand, a network formed at the electroweak scale can produce vortons accounting for up to $6\\%$ of the critical density. Some consequences of our results are discussed. ", "introduction": "\\label{v-in} As first pointed out by Witten \\cite{witten}, cosmic strings can in some circumstances (typically when the electromagnetic gauge invariance is broken inside the string) behave as `superconducting wires' carrying large currents and charges---up to the order of the string mass scale in appropriate units. The charge carriers can be either bosons or fermions (see \\cite{vs} for a review). The former type occurs when it becomes energetically favourable for a charged Higgs field to have a non-zero vacuum expectation value in the string core; the latter happens when fermions couple to the string fields creating fermion zero modes. It is well known that arbitrarily large currents are not allowed---there is a critical value beyond which the current saturates. In other words, for large enough winding number per unit length, the superconducting condensate is quenched down, suppressing the current flow. Also, the current can decay by magnetic flux-line tunnelling; this can be used to impose constraints on allowed particle physics models. If superconducting strings carry currents, they must also carry charges of similar magnitude. This includes not only charges trapped at formation by the Kibble mechanism but also the ones due to string inter-commuting between regions of the string network with different currents. Just like with currents, charge densities cannot have arbitrarily large magnitude---there is a limit beyond which there will no longer be an energy barrier preventing the charge carriers from leaving the string. A rather important point is that the presence of charges on the string tends to counteract the current quenching effect discussed above. In fact, numerical simulations of contracting string loops at fixed charge and winding number have shown \\cite{davsh} that a `chiral' state with equal charge and current densities is approached as the loop contracts. In this limiting chiral case, quenching is in fact eliminated completely. This has several important consequences. Strings that have trapped charges as a consequence of a phase transition can become superconducting even if the formation of a condensate was otherwise energetically unfavoured. More importantly, a string with both a charge and a current density will have a non-zero angular momentum. In the cosmological context, these strings would of course interact with the cosmic plasma, originating a number of interesting consequences. The most remarkable of these, however, has to do with the evolution of string loops. If a superconducting string loop has an angular momentum, it is semi-classically conserved, and it tries to resist the loop's tension. This will at least increase the loop's lifetime. If the current is too large, charge carriers will leave the string accompanied by a burst of electromagnetic radiation, but otherwise it is possible that dynamically stable loops form. These are called vortons \\cite{vor}---they are stationary rings that do not radiate classically, and at large distances they look like point particles with quantised charge and angular momentum. Their cosmological significance comes from the fact that they provide very strong constraints on allowed particle physics models, since they behave like non-relativistic particles. According to current belief \\cite{vor,vor2}, if they are formed at high enough energy scales they are as dangerous as magnetic monopoles, producing an over-density of matter in disagreement with observations. On the other hand, low-mass vortons could be a very interesting dark matter candidate. Understanding the mechanisms behind formation and evolution is therefore an essential cosmological task. The overwhelming majority of the work done on cosmic strings so far was concerned with the structureless Goto-Nambu strings (but see \\cite{carternew} and references therein for some exceptions). In the case of work on vortons, this means that somewhat {\\em ad-hoc} estimates had to be made for some properties of the cosmic string network---notably for microscopic quantities such as current and charge densities. This is despite the fact it has been recognised a long time ago that, even though they might be computationally very useful \\cite{ms,ms2,ms2a}, Goto-Nambu models cannot realistically be expected to account for a number of cosmologically relevant phenomena, due to the very limited number of degrees of freedom available. Two such phenomena are the build-up of small-scale structure and charge and current densities. In this paper we fill this important gap by discussing the problem of vorton formation in the context of the superconducting string models of Witten \\cite{witten} and of Carter and Peter \\cite{wcp} (sections \\ref{v-wt} and \\ref{v-cp}). Strangely enough, the issue of the conditions for vorton formation has been so far neglected with respect to those of their stability and cosmological consequences. We will start by introducing these models and determining the microscopic string equations of motion in each case. It will be shown that in the relevant chiral limit these equations coincide---this also provides the first conclusive evidence of the validity of the supersonic elastic models of Carter and Peter \\cite{wcp}. We then proceed to study the evolution of a number of loop solutions of these equations numerically (sections \\ref{v-fl} and \\ref{v-ex}), and from the results of this analysis parameters will be introduced which characterise the loop's ability to evolve into a vorton state (section \\ref{v-vt}). Finally, we discuss a very simple phenomenological model for the evolution of the superconducting currents on the long cosmic string network \\cite{mss}, based on the dynamics of a `superconducting correlation length' (sections \\ref{v-cr}--\\ref{v-ff}). Using this model we can therefore estimate the currents carried by string loops formed at all relevant times, and thus (in principle) decide if these can become vortons (section \\ref{v-rs}) and calculate the corresponding density (section \\ref{v-den}). Based on our results, we don't expect any GUT vortons to form at all. This is essentially because the friction-dominated epoch is very short for GUT-scale strings \\cite{ms}, so their currents and charges are never large enough to prevent them from becoming relativistic---and therefore liable to losses. Even if they did form, they wouldn't be in conflict with the standard cosmological scenario if they decayed soon after the end of the friction-domination epoch. Hence we conclude that, in contrast with previously existing estimates \\cite{vor,vor2}, one cannot at the moment rule out GUT superconducting string models. We should point out at the outset that there are essentially three improvements in the present work which justify the different end result for GUT-scale strings. Firstly, by analysing simple (but physically relevant) loop solutions of the microscopic string equations of motion for the Witten model, we can get a much improved idea of how superconducting loops evolve and of how (and under which conditions) they reach a vorton state. Secondly, by using a simple model for the evolution of the currents on the long strings \\cite{mss} we can accurately determine the typical currents on each string loop at the epoch of its formation. Finally, the use of the analytic formalism previously introduced by the present authors \\cite{ms,ms2a} allows us to use a quantitative description throughout the paper, and in particular to determine the loop sizes at formation. As will become clear below, when taken together these allow a detailed analysis of the process of vorton formation to be carried out, either in the Witten model (as is done in this paper) or any other that one considers relevant. In contrast, note that Davis \\& Shellard \\cite{vor} restrict themselves to the particular case of the initial Brownian Vachaspati-Vilenkin loops with Kibble currents, and do not consider the subsequent evolution of the network. On the other hand, Brandenberger {\\em et al.} \\cite{vor2} make rather optimistic order-of-magnitude estimates about the process of relaxation into a vorton state. As it turns out, for high energy GUT scales, all these loops become relativistic before reaching a vorton state. Finally, neither of these treatments has the benefit of a quantitative model for the evolution of the long-string network \\cite{ms} which allows one to accurately describe the process of loop production. On the other hand, as we lower the string-forming energy scale we expect more and more efficient vorton production, and the 'old' scenario still holds. Therefore intermediate-scale superconducting strings are still ruled out, since they would lead to a universe becoming matter-dominated earlier than observationally allowed. Finally, at low enough energy scales, vortons will be a dark matter candidate. For example, for a string network formed around $T\\sim10^2\\,GeV$ (typical of the electroweak phase transition) they can provide up to $6\\%$ of the critical density. A more detailed discussion of these issues is left to a forthcoming publication \\cite{inprep}. Throughout this paper we will use fundamental units in which $\\hbar=c=k_B=Gm^2_{Pl}=1$. ", "conclusions": "\\label{v-cc} In this paper we have presented the first rigourous study of the cosmological evolution of superconducting strings in the limit of chiral currents. We have shown that in this limit the elastic string model of Carter \\& Peter\\cite{wcp} coincides with the model derived from first principles by Witten\\cite{witten}. By analysing physically relevant loop solutions of the microscopic equations of motion for these strings, we have verified that the effect of frictional damping is crucial for vorton formation. We then defined suitable parameters characterising the evolution of these loops, and in particular whether or not they become vortons. In particular, we have established the usefulness of the `stability parameter' ${\\overline n}$. In general, it is more difficult to form vortons when the string-forming phase transition is of first order. This is because such networks produce, during their evolution in the stretching regime, loops with a size close to that of the horizon; these will therefore be significantly affected by expansion, which tends to decrease the fraction of the loops's energy in the current---whereas friction tends to increase it. After introducing a simple `toy model' for the evolution of currents on the strings \\cite{mss}, we have considered the cases of first and second-order GUT-scale string-forming and superconducting phase transitions (which is the most favourable GUT case of vorton formation since frictional forces can act longer). We have presented evidence suggesting that GUT-scale string networks might well produce no vortons, and that even if they do, this will not necessarily rule out such models. This is in contradiction with previous, less detailed studies \\cite{vor,vor2}, and hence calls for a re-examination of a number of cosmological scenarios involving superconducting strings. Notably, these strings could be at the origin of the observed galactic magnetic fields \\cite{msmag}. Finally, we have explicitly calculated the vorton density in two `extreme' cases to illustrate the method that one should follow once the microphysical properties of these networks are known in more detail. For electroweak-scale string networks, we have found that vortons can produce up to about $6\\%$ of the critical density of the universe. On the other hand, it is conceivable that superconducting string networks formed at an energy scale $T\\sim10^4-10^6\\,GeV$ (depending on details of the model) can solve the dark matter problem. The detailed analysis presented in this paper for GUT stings can obviously be extended to other energy scales---this will be the subject of a forthcoming publication \\cite{inprep}. Obviously, as we lower the energy scale, the frictional force becomes more and more important and acts for a longer time. Hence the vorton-forming region of parameter space increases, and by the electroweak scale almost all loops chopped off the long-string network will become vortons. We therefore conclude that in addition to the low-$G\\mu$ regime (which as we saw includes the electroweak scale) where vortons can be a source of dark matter and to an intermediate-$G\\mu$ range in which vortons would be too massive to be compatible with standard cosmology (thereby excluding these models), there is also a high-$G\\mu$ regime (of which the GUT scale is part) in which vortons don't form at all and therefore no cosmological constraints based on them can be set. It is then curious (to say the least) that vorton constraints can be used to rule out cosmic string models in a wide range of energy scales $G\\mu$, but not those formed around the GUT or the electroweak scales, where cosmic strings can be cosmologically useful." }, "9804/astro-ph9804235_arXiv.txt": { "abstract": "We investigate the possibility of accounting for the currently inferred primordial abundances of D, $^{3}$He, $^{4}$He, and $^{7}$Li by big bang nucleosynthesis in the presence of baryon density inhomogeneities plus the effects of late--decaying massive particles (X), and we explore the allowed range of baryonic fraction of the closure density $\\Omega_{b}$ in such context. We find that, depending on the parameters of this composite model (characteristic size and density contrast of the inhomogeneities; mass--density, lifetime, and effective baryon number in the decay of the X particles), values as high as $\\Omega_{b}h_{50}^{2}\\simeq 0.25-0.35$ could be compatible with the primordial abundances of the light nuclides. We include diffusion of neutrons and protons at all stages, and we consider the contribution of the X particles to the energy density, the entropy production by their decay, the possibility that the X--products could photodissociate the light nuclei produced during the previous stages of nucleosynthesis, and also the possibility that the decay products of the X--particles would include a substantial fraction of hadrons. Specific predictions for the primordial abundance of Be are made. ", "introduction": "Standard homogeneous big bang nucleosynthesis could have produced the observationally inferred primordial abundances of D, $^{3}$He, $^{4}$He, and $^{7}$Li, provided that the baryon fraction of the cosmic closure density $\\Omega_{b}$ would lie in the range: $$0.04\\lapprox \\Omega_{b}h_{50}^{2}\\lapprox 0.08\\eqno(1)$$ \\noindent where $h_{50}$ is the Hubble constant in units of 50 km s$^{-1}$ Mpc$^{-1}$ (Walker et al. 1991; Copi, Schramm, \\& Turner 1995). For the long--time most favored cosmological model, a flat Universe with $\\Omega_{M} = 1$ and $\\Omega_{\\Lambda} = 0$ (those being, respectively, the fractional contributions of matter and vacuum energy densities to the closure density), the upper bound to $\\Omega_{b}$ would mean that most matter in the Universe should be in nonbaryonic form. Given the far--reaching implications of the dominance of nonbaryonic dark matter, possible alternatives to homogeneous big bang nucleosynthesis have been explored, especially during the last 15 years or so. The suggestion that the quark--hadron phase transition might be first--order and generate baryon inhomogeneities (Witten 1984) led to the calculation of the possible effects on primordial nucleosynthesis (Applegate \\& Hogan 1985; Applegate, Hogan, \\& Scherrer 1987; Malaney \\& Fowler 1988). The goal was to see whether inhomogeneous big bang nucleosynthesis with $\\Omega_{b} = 1$ might account for the primordial light--element abundances. Besides a first--order quark--hadron phase transition, other mechanisms might also generate baryon inhomogeneities. Much of the work in this line is reviewed by Malaney \\& Mathews (1993). However, the recent studies, treating accurately the coupling between baryon diffusion and nucleosynthesis, show that the upper limit on $\\Omega_{b}$ set by the light--element abundances does not significantly differ from that obtained for homogeneous big bang nucleosynthesis (Mathews, Schramm, \\& Meyer 1993; Thomas et al. 1994). This last conclusion, though, has very recently been challenged by Orito et al. (1997), who explore the dependence of primordial nucleosynthesis on the geometry of baryon inhomogeneities and find that cylindrical geometry might allow to satisfy the observational constraints with baryon fractions as high as $\\Omega_{b}h_{50}^{2}\\lapprox 0.2$. A different approach has been to explore the possible modifications of the yields from homogeneous big bang nucleosynthesis by the effects of the decay of unstable massive particles ($M\\gapprox few\\ GeV$), produced at earlier stages in the evolution of the Universe and with half--lives longer than the standard nucleosynthesis epoch ($\\tau_{x}\\gapprox 10^{4}\\ s$) (Audouze, Lindley, \\& Silk 1985; Dom\\'{\\i}nguez--Tenreiro 1987; Yepes \\& Dom\\'{\\i}nguez--Tenreiro 1988; Dimopoulos et al. 1988). Gravitinos produced during reheating at the end of inflation are a possible example of such particles. In Dimopoulos et al. (1988), the emphasis is put on the resulting hadron cascade. The main problem encountered in this model is the predicted overproduction of $^{6}$Li: $^{6}Li/^{7}Li\\gg 1$, whereas observations show $^{6}Li/^{7}Li\\lapprox 0.1$. Although they have only been considered separately, baryon inhomogeneities and the presence of unstable massive particles decaying when the Universe has already cooled down below $T_{9}\\simeq 0.4$ are by no means mutually exclusive. Here we explore their combined effects on the primordial abundances of the light elements. The parameter space now has, of course, a dimension which is the sum of those for the two separate cases: characteristic size and density contrast of the inhomogeneities, mass--density, lifetime, and mode of decay of the massive particles. We find that there are regions in such extended parameter space where values of $\\Omega_{b}$ as high as $\\Omega_{b}h_{50}^{2}\\simeq 0.35$ would still be compatible with the primordial abundances of the light nuclides inferred from observations. Such values of $\\Omega_{b}$ are of the same order as the low values for $\\Omega_{M}$ currently derived from a variety of sources, including high--redshift supernova searches (Perlmutter et al. 1998; Garnavich et al. 1998). Our results thus suggest again the possibility that all the matter in the Universe could be baryonic. On the other hand, recent determinations of the D abundance in high--redshift QSO absorbers, when confronted with the currently inferred primordial $^{4}$He abundance, might be in conflict with the predictions of standard, homogeneous big bang nucleosynthesis for $N_{\\nu} = 3$ (Steigman 1998): the ``low'' high--redshift D abundances (which appear more reliable) would indicate too high a value of $\\Omega_{b}$ to be compatible with that corresponding to the $^{4}$He abundance. Since it is hard to tell whether this conflict points to new physics or just to systematic errors in the derivation of abundances, Steigman, Hata, \\& Felten (1998) have discarded the constraint on $\\Omega_{b}$ from standard big bang nucleosynthesis and turned to other observational constraints to determine the key cosmological parameters. The results from our composite model, by showing how minor deviations from the standard hypotheses can produce agreement with the primordial abundances inferred from observations, support that attitude. Besides, as we will see, the combined effects of inhomogeneities plus late--decaying particles might solve the conflict between D and $^{4}$He abundances. ", "conclusions": "We have shown, by means of a simple model, that the combined effects on big bang nucleosynthesis of baryon inhomogeneities plus the decay of unstable, relatively long--lived massive particles, giving rise to both electromagnetic and hadron cascades, might be to allow agreement with the primordial light--element abundances inferred from observations for values of $\\Omega_{b}$ much higher that those allowed by standard, homogeneous nucleosynthesis. The upper limit might be as high as $\\Omega_{b}h_{50}^{2}\\simeq 0.35$. The values obtained here are of the same order as the low $\\Omega_{M}$ values now being derived from a variety of sources and, therefore, they pose in new terms the question of whether all matter in the Universe could be baryonic. A testable prediction of the model is the production of a $^{9}$Be abundance that is of the order of current observational upper limits. On the other hand, in the parameter region of our model where there is agreement between predicted and observationally inferred primordial light--element abundances, given values of $\\Omega_{b}$ (or, equivalently, $\\eta_{10}$) always predict ``low'' D abundances (in the sense of the high--redshift abundances referred to in the Introduction), thus potentially eliminating the conflict with the $^{4}$He abundance pointed out by Steigman (1998). The model presented here deals with inhomogeneities in a very simplified way. A futher step will be to examine the effects of the geometry of the density fluctuations on the outcome. Orito et al. (1997) have already shown that cylindrical shell geometry alone (without the extra effects of late--decaying particles) might allow $\\Omega_{b}\\lapprox 0.2$ (but for density contrasts $R\\sim 10^{6}$, much higher than those considered here). Another extension of the model will be to consider particles with shorter half--lives, decaying at the time when thermonuclear reactions are still taking place." }, "9804/astro-ph9804145_arXiv.txt": { "abstract": "The neutrino emissivities resulting from direct URCA processes in neutron stars are calculated in a relativistic Dirac-Hartree approach in presence of a magnetic field. In a quark or a hyperon matter environment, the emissivity due to nucleon direct URCA processes is suppressed relative to that from pure nuclear matter. In all the cases studied, the magnetic field enhances emissivity compared to the field-free cases. ", "introduction": " ", "conclusions": "" }, "9804/astro-ph9804286_arXiv.txt": { "abstract": "We present a detailed morphological analysis of the galaxy populations in the first two clusters to be completed in an extensive observational study of nine high-redshift clusters of galaxies (Oke, Postman \\& Lubin 1988). These two clusters, CL0023+0423 and CL1604+4304, are at redshifts of $z = 0.84$ and $z = 0.90$, respectively. The morphological studies are based on high-angular resolution imagery taken with WFPC2 aboard the {\\it Hubble Space Telescope}. These data are combined with deep, ground-based $BVRI$ photometry and spectra taken with the Keck 10-meter telescopes. The morphological classifications presented in this paper consist of two parts. Firstly, we provide a quantitative description of the structural properties of $\\sim 600$ galaxies per cluster field using the Medium Deep Survey automated data reduction and object classification software (Griffiths et al.\\ 1994; Ratnatunga, Ostrander \\& Griffiths 1997). This analysis includes the galaxy position, photometry, and best-fit bulge+disk model. Secondly, for the brightest subsample of $\\sim 200$ galaxies per cluster field, we provide a more detailed morphological description through a visual classification based on the revised Hubble classification scheme (e.g.\\ Sandage 1961; Sandage \\& Bedke 1994). Based on these classifications, we have examined the general relation between galaxy morphology and other photometric and spectral properties. We find that, as expected, the elliptical and S0 galaxies are redder, on average, than the spirals and irregulars. In addition, there is a strong correlation between morphology and spectral type. Of the galaxies that are visually classified as ellipticals, the majority show K star absorption spectra which are typical of nearby, red early-type galaxies; however, a few are actually blue compact galaxies with spectra characterized by fairly strong, narrow emission lines. Normal late-type galaxies typically have spectra with blue colors and [\\ion{O}{2}] emission, while the presence of strong star-formation features, such as extremely high equivalent width [\\ion{O}{2}], ${\\rm H\\beta}$, and/or [\\ion{O}{3}] emission, is always accompanied by peculiar morphologies which suggest recent mergers or interactions. We have used the statistical distributions of cluster galaxy morphologies to probe the overall morphological composition of these two systems. This analysis reveals that the two clusters contain very different galaxy populations. CL0023+0423 has a galaxy population which is more similar to groups of galaxies and the field. This system is almost completely dominated by spiral galaxies. CL1604+4304, however, has a morphological composition which is more typical of a normal, present-day cluster; early-type galaxies comprise $\\sim 76\\%$ of all galaxies brighter than $M_{V} = -19.0 + 5~{\\rm log}~h$ in the central $\\sim 0.5~h^{-1}~{\\rm Mpc}$. The ratio of S0 galaxies to ellipticals in this cluster is $1.7 \\pm 0.9$, consistent with local cluster populations. The morphological results support the conclusions of the dynamical analysis presented in the second paper of this series (Postman, Lubin \\& Oke 1998). CL0023+0423 consists of two galaxy groups which are separated by $\\sim 2900~{\\rm km~s^{-1}}$ in radial velocity. CL1604+4304, on the other hand, has a velocity distribution indicating that it is already well-formed and relaxed. The morphological composition, velocity dispersion, and implied mass of the CL1604+4304 system are consistent with an Abell richness class 2 or 3 cluster. ", "introduction": "The study of the galaxy populations of rich clusters provides important constraints on the formation mechanisms of both clusters and galaxies. Present--day clusters show a distinct correlation between the structure of the cluster and the galaxy population. Irregular, open clusters, such as Virgo, are spiral--rich. These systems show no single, central condensation, though the galaxy surface density is at least five times as great as the surrounding field ($n_{\\rm gal} > 30~h^{3}~{\\rm galaxies~{Mpc}^{-3}}$). These clusters are often highly asymmetric and have significant degrees of substructure. Dense, centrally concentrated clusters, such as Coma, contain predominantly early--type galaxies in their cores (Abell 1958; Oemler 1974; Dressler 1980a,b; Postman \\& Geller 1984). These clusters have a single, prominent concentration among the bright member galaxies and typically display a high--degree of spherical symmetry, though this does not preclude evidence of some substructure. Central densities can reach as high as $10^{4}~h^{3}~{\\rm galaxies~{Mpc}^{-3}}$ (e.g\\ Bahcall 1975; Dressler 1978). In these regions, spiral galaxies comprise less than 10\\% of the cluster population, while elliptical (E) and S0 galaxies make up 90\\% or more of the population. The ratio of S0s to ellipticals is typically S0/E $\\sim 2$ (Dressler 1980a). The galaxy content of clusters is part of the general morphology--density relation of galaxies; as the local density increases, the fraction of elliptical and S0 galaxies increases, while the fraction of spiral galaxies decreases (Hubble 1936; Dressler 1980a,b; Postman \\& Geller 1984). Previous studies of clusters of galaxies at $z < 1$ have revealed significant evolution in the morphology and the color of the cluster members. One of the most notable of these changes is the progressive blueing of cluster's galaxy population with redshift, a trend first observed by Butcher \\& Oemler (1984). They found that the fraction of blue galaxies in a cluster is an increasing function of redshift, indicating that clusters at redshifts of $z \\sim 0.5$ are significantly bluer than their low--redshift counterparts. At redshifts of $z \\sim 0.4$, the fraction of blue galaxies is $\\sim 20\\%$. Recent HST image data reveal that many of these blue galaxies are either ``normal'' spirals or have peculiar morphologies, resulting in non--elliptical fractions which are 3 to 5 times higher than the average current epoch cluster (Dressler et al.\\ 1994; Couch et al.\\ 1994; Oemler, Dressler \\& Butcher 1997; Dressler et al.\\ 1997). Detailed photometric observations of other intermediate redshift ($z \\simless 0.4$) clusters have confirmed the original results of Butcher \\& Oemler (e.g.\\ Millington \\& Peach 1990; Luppino et al.\\ 1991; Rakos \\& Schombert 1995). Even though these clusters show an increased fraction of blue galaxies, they still contain a population of E/S0s distinguished by extremely red colors and a tight color--magnitude (CM) relation (a ``red envelope''). Both the mean color and the CM relation are consistent with that of present--day ellipticals (e.g.\\ Sandage 1972; Butcher \\& Oemler 1984; Arag\\'on-Salamanca et al.\\ 1991; Luppino et al.\\ 1991; Molinari et al.\\ 1994; Dressler et al.\\ 1994; Smail, Ellis \\& Fitchett 1994; Stanford, Eisenhardt \\& Dickinson 1995). At redshifts of $z \\simgreat 0.4$, the red envelope has moved bluewards with redshift (Arag\\'on-Salamanca et al.\\ 1993; Smail et al.\\ 1994; Rakos \\& Schombert 1995; Oke, Gunn \\& Hoessel 1996; Lubin 1996; Ellis et al.\\ 1997; Stanford, Eisenhardt \\& Dickinson 1997). At $z \\sim 0.9$, there are few cluster members with colors nearly as red as present--day ellipticals. The color distribution of this high-redshift elliptical population is relatively narrow, and the trend is uniform from cluster to cluster; this suggests a homogeneous population which formed within a narrow time span (e.g.\\ Bower, Lucey \\& Ellis 1992a,b). Dickinson (1995) finds similar results in a cluster of galaxies which is associated with the $z = 1.206$ radio galaxy 3C 324. The galaxies in this cluster exhibit a narrow, red locus in the CM magnitude diagram. This branch is $\\sim 0.6$ mag bluer than the expected ``no--evolution'' value, though the intrinsic rms color scatter is only 0.2 mag. The observed color trend for the red envelope of galaxies in this data is consistent with passive evolution of an old stellar population formed by a single burst of star formation at redshifts of $z \\simgreat 2$. The reasonably small color scatter would imply closely synchronized intra--cluster star formation (Bower et al.\\ 1992a,b; Arag\\'on-Salamanca et al.\\ 1993; Dickinson 1995; Ellis et al.\\ 1997; Stanford et al.\\ 1997). The high-resolution imaging of HST has been essential in understanding the evolutionary processes occurring at intermediate redshifts (see e.g.\\ Abraham et al.\\ 1996a). Morphological classifications can be made on scales of $\\sim 1~{\\rm kpc}$, providing a direct comparison with ground-based classifications of nearby galaxies. A comprehensive survey of 10 clusters of galaxies at $z = 0.37 - 0.56$ has revealed a significant change relative to local clusters in the composition and behavior of the galaxy populations (Smail et al.\\ 1997, hereafter S97; Dressler et al.\\ 1997, hereafter D97). The authors have visually classified over 6000 galaxies based on the Revised Hubble Scheme used to classify nearby galaxies (e.g.\\ Sandage 1961; Sandage \\& Bedke 1994). These classifications are used to quantify the morphological composition of each cluster. Their results indicate that the morphology--density relation is qualitatively similar to that in the local universe in those intermediate redshift clusters which are centrally-concentrated and compact; however, the relation is non-existent in the loose, open clusters. Even so, all of the clusters exhibit a roughly similar make-up of galaxy morphologies. The fraction of ellipticals is the same or larger than that in local clusters; the S0 fraction, however, is $\\sim 2-3$ times lower, with a corresponding increase in the cluster spiral population. These findings imply that the elliptical population is already in place by $z \\sim 0.5$, but a large fraction of the S0 galaxies are formed between redshifts of $z \\sim 0.5$ and $z = 0$ (D97). However, it should be noted that these classifications are typically derived from images which are not of comparable quality to the local data. Because of such uncertainty, the observed evolution in the S0 population is still in contention (e.g.\\ Stanford et al.\\ 1997; Andreon, Davoust \\& Helm 1997; Andreon 1998). Because there appears to be significant evolution occurring between redshifts of $z \\sim 0.5$ and the present epoch, it is critical to extend these detailed observations to even higher redshifts if we are to understand the formation of galaxy morphology, as well as the mechanisms and timescales of this evolution. Therefore, we have undertaken an extensive observational program to study nine candidate clusters of galaxies at redshifts of $z > 0.6$. The cluster sample was chosen from the Gunn, Hoessel \\& Oke (1986) survey and the Palomar Distant Cluster Survey (PDCS; Postman et al.\\ 1996). For each cluster, we are in the process of obtaining deep $BVRI$ photometry from Keck and deep $K$ photometry from the KPNO 4-meter, low-resolution spectra from Keck, and high angular resolution imagery from HST. The observations and data processing procedures of this survey are the subject of the first paper in this series (Oke, Postman \\& Lubin 1998; hereafter Paper I). The first two clusters to be completed in this observational program are CL0023+0423 and CL1604+4304 at redshifts of $z = 0.84$ and $z = 0.90$, respectively (see Paper I). In this paper, we have used HST images to undertake a detailed morphological analysis of the galaxy populations in the central regions of these two clusters. The reduction and analysis of the Keck $BVRI$ photometry and spectra of the galaxies in these cluster fields are discussed in the second paper of this series (Postman, Lubin \\& Oke 1998; hereafter Paper II). However, the galaxy parameters presented in Paper II are used in this paper, specifically for a comparison with the morphological properties. In Sect.\\ 2, we provide a brief description of the data. In Sects.\\ 3 and 4, we describe the automated and visual galaxy classification procedures used in this paper and present a comparison between the two techniques. In Sect.\\ 5, we examine the morphologies of the galaxies in the two cluster fields. This includes the relationship between morphology and other galaxy properties, as well as the overall distribution of morphologies in the cluster. A discussion and summary of our conclusions are presented in Sects.\\ 6 and 7. In the following analyses, we have assumed $q_{0} = 0.1$ (e.g\\ Carlberg et al.\\ 1996) and $H_{0} = 100~h~{\\rm km~s^{-1}~Mpc^{-1}}$. ", "conclusions": "As part of an observational program to study distant clusters of galaxies, we have examined the morphological properties of the galaxies in two cluster fields, CL0023+0423 at $z = 0.84$ and CL1604+4304 at $z = 0.90$, using high-resolution HST images. The morphology of the individual galaxies have been studied by two methods; 1) a quantitative description of the structural properties of $\\sim 600$ galaxies per cluster field is provided by the Medium Deep Survey automated data reduction and ``bulge+disk'' object classification software; 2) the brightest subsample of $\\sim 200$ galaxies per cluster field are assigned a more detailed morphological description through a visual classification based on the revised Hubble scheme. A comparison between the two techniques shows that there is a reasonable correlation between the parameters of the automated and visual classifications (see also Lahav et al.\\ 1995). To investigate the morphological composition of the two galaxy clusters, we have used the visual classifications of the brightest subsample of galaxies in each field. Our main conclusions are summarized below. \\newcounter{discnt} \\begin{list} {\\arabic{discnt}.} {\\usecounter{discnt}} \\item The color-magnitude diagrams and the color histograms of all (field + cluster) galaxies in the two cluster field show a clear progression in color between early- and late-type galaxies. As expected, the elliptical and S0 galaxies are redder, on average, than the spirals and irregulars. This trend is also apparent in the color ages which represent the time since the last period of major star formation. 83\\% of the galaxies classified as late-type (spiral or irregular/peculiar) have color ages less than 2 Gyr. In contrast, 55\\% of the galaxies classified as early-type have color ages of greater than 2 Gyr, and 73\\% of all galaxies with ages greater than 3 Gyr are classified as early-type. In addition, cluster galaxies are typically older than field galaxies at similar redshifts. This is due in large part to the fact that there are more early-type galaxies in these systems. \\item We see a distinct correlation between the galaxy morphology and the corresponding spectral features. We have specifically examined this relation for those galaxies which are confirmed cluster members. The majority of galaxies that are visually classified as ellipticals show spectra which are typical of nearby, red elliptical galaxies. However, some of the galaxies visually classified as ellipticals turn out to be blue compact galaxies with spectra characterized by fairly strong, narrow emission lines. Normal late-type galaxies typically have spectra with blue colors and [\\ion{O}{2}] emission, while the presence of strong star-formation features, such as extremely high equivalent width [\\ion{O}{2}], ${\\rm H\\beta}$, and/or [\\ion{O}{3}] emission, is always accompanied by peculiar morphologies which suggest recent mergers or interactions. \\item Despite being at very similar redshifts, the two cluster systems contain very different galaxy populations as indicated by their background-subtracted morphological distributions. We have examined all galaxies brighter than $M_{V} = -19.0 + 5~{\\rm log}~h$ in the central $\\sim 0.5~h^{-1}~{\\rm Mpc}$ of the cluster. CL0023+0423 has a galaxy population which is more typical of groups and the field population. The numbers from the statistical distribution are consistent with almost all of the galaxies being normal spirals. The spectral results support these numbers, independently indicating spiral fractions of 66\\% or more. CL1604+4304, in contrast, has a morphological composition which is characteristic of a normal, present-day rich cluster. Early-type galaxies comprise 76\\% of all galaxies in this region. In this population, the ratio of S0 galaxies to ellipticals is $1.7 \\pm 0.9$, consistent with galaxy populations found in local clusters (Dressler 1980a). \\item The morphological results support the conclusions of the dynamical analysis presented in Paper II. CL0023+0423 is apparently two independent systems separated in velocity by $\\sim 2900~{\\rm km~s^{-1}}$. The velocity dispersions and implied masses indicate that these systems are similar to local galaxy groups or poor clusters. Though this may be a chance projection, the dynamical and morphological evidence may indicate that we are seeing the merger of two spiral-dominated galaxy groups (see Lubin, Postman \\& Oke 1998a). The velocity histogram of CL1604+4304, on the other hand, is consistent with a Gaussian distribution, implying that this system formed at an earlier epoch and is already relaxed. The velocity dispersion and implied mass of this system are consistent with an Abell richness class 2 or 3 cluster. \\end{list} \\vskip 0.5cm We thank the anonymous referee for his thorough review of this paper. Alan Dressler, Chris Fassnacht, and Ian Smail are thanked for useful discussions, comments, and material aids to this paper. It is also a great pleasure to thank Allan Sandage for his generous gift of time and invaluable expertise to this project. The W.M. Keck Observatory is operated as a scientific partnership between the California Institute of Technology, the University of California, and the National Aeronautics and Space Administration. It was made possible by generous financial support of the W. M. Keck Foundation. LML graciously acknowledges support from a Carnegie Fellowship. This research was supported in part by {\\it HST} GO analysis funds provided through STScI grant GO-06000.01-94A and {\\it HST} Archival grant 7536. \\clearpage" }, "9804/astro-ph9804208_arXiv.txt": { "abstract": "s{Galactic synchrotron emission is a potentially confusing foreground, both in total power and in polarization, to the Cosmic Microwave Background Radiation. It also contains much physical information in its own right. This review examines the amplitude, angular power spectrum and frequency spectrum of the synchrotron emission as derived from the presently available de-striped maps. There are as yet no maps at arcminute resolution at frequencies above 2.4 GHz. This incomplete information is supplemented with data from supernovae, which are thought to be the progenitors of the loops and spurs found in the Galactic emission. The possible variations of the frequency spectral index from pixel to pixel are highlighted. The relative contributions of free-free and synchrotron radiation are compared, and it is concluded that the free-free contribution may be smaller than had been predicted by COBE. New high resolution polarization surveys of the Galactic plane suggest detail on all scales so far observed. At high latitudes the large percentage polarisation means that the foreground contamination of the polarised CMB signal will be more serious than for the unpolarized radiation.} ", "introduction": "Galactic emission at radio wavelengths is important to understand in its own right. Moreover it is crucial to be able to quantify and remove this component as a foreground to the cosmic microwave background (CMB). Both synchrotron and free-free emission contribute to this foreground, with the synchrotron emission dominating at low frequencies ($\\leq$1 GHz). The synchrotron emissivity is a function of both the relativistic (cosmic ray) density and the local magnetic field strength. The luminosity at frequency $\\nu$ is given by \\be I(\\nu) = const \\; L N_0 B^{(p+1)/2} \\nu^{-(p-1)/2} \\ee \\noindent where $N_0$ is the density of relativistic electrons, $L$ is the emission depth, $B$ is the magnetic field and the relativistic electron energy spectrum\\cite{ve} is given by $dN/dE = N_0 E^{-p}$. The radio spectral index is $\\alpha= (p-1)/2$ in energy terms or $2+\\alpha$ when expressed as a brightness temperature $T_B \\propto \\nu^{-(2+\\alpha)} = \\nu^{-\\beta}$. Within the interstellar magnetic field of 2 to 5 microgauss, emission at GHz frequencies is characteristically from relativistic electrons with an energy of 1 to 10 GeV. Both $B$ and $N_0$, as well as $p$, will vary from point to point in the Galactic disk and nearby halo. The cosmic ray electrons are thought to originate mainly in supernovae then diffuse outwards in the expanding remnant. Structure will be formed in the remnant as it collides with the non-uniform ambient medium. The magnetic field will be likewise amplified in compression regions and vary in strength and direction. The net effect is to produce elongated synchrotron emission structures on a wide range of scales. The spectral index of the emission will vary with position for two reasons. Firstly the electron spectral index varies from one supernova to another and secondly the spectrum steepens ($\\Delta p = +1$) with time due to radiation energy loss thus giving an age-dependent spectral index. This paper will describe the synchrotron features in and near the Galactic plane which are believed to give rise to the structures seen at higher galactic latitudes. All-sky and large area surveys are assessed to give information about the amplitude and spectrum of the high latitude emission which is a potential confusing foreground to the CMB. Comments are given about the role of synchrotron polarization and of free-free emission. ", "conclusions": "" }, "9804/astro-ph9804102_arXiv.txt": { "abstract": "We report on {\\it ROSAT} HRI observations of the $z=0.61$ radio galaxy 3C\\,220.1. The X-ray emission from this object consists of an extended component, which we attribute to luminous cluster emission, and a compact central source. The compact component is too bright to be modelled as a cooling flow under some plausible assumptions for the hot gas temperature and distribution; we suggest instead that it is directly related to the core of the radio source. The X-ray flux of the compact component is consistent with the prediction of Worrall \\etal\\ (1994) that all powerful radio galaxies should have a central jet-related X-ray emission component that is proportional in strength to the radio core flux density. Other observations of distant 3CR radio sources are consistent with this model. ", "introduction": "In unified models of powerful radio sources (e.g.\\ Barthel 1989) core-dominated quasars, lobe-dominated quasars and radio galaxies are the same objects, with the apparent differences attributed to the effects of relativistic beaming and anisotropic obscuration on a population of sources oriented randomly with respect to the line of sight. In quasars we see the AGN directly, while radio galaxies have their jet axis at large angles to the line of sight so that a torus of gas and dust obscures optical continuum and broad-line emission from the AGN. This torus should also obscure soft X-ray emission originating close to the central engines, leading to suggestions (e.g.\\ Crawford \\& Fabian 1996) that X-ray emission from powerful radio galaxies should in general be dominated by thermal emission from hot cluster gas. However, an additional X-ray component may arise from the radio-emitting plasma directly, through synchrotron or synchrotron self-Compton radiation, or indirectly, through mechanisms such as the inverse-Compton scattering of external photons. If the different components can be separated, X-ray observations can provide physical insights into the active nucleus, jets and large-scale environment. There is substantial evidence that unabsorbed radio-related non-thermal X-rays are seen. Fabbiano \\etal\\ (1984) first stressed that a correlation between the total soft X-ray and radio-core luminosity in radio galaxies implied a nuclear, jet-related origin for at least some of the X-ray emission. More recent work has strengthened this conclusion, as high-resolution X-ray observations have allowed point-like and extended components to be separated (Worrall \\& Birkinshaw 1994; Edge \\& R\\\"ottgering 1995; Worrall 1997). Although component separation is generally better for closer, less powerful radio galaxies, compact soft X-ray emission is also seen in powerful narrow-line FRII radio galaxies that are known to lie in sparse environments, where cluster emission is not a source of confusion (e.g.\\ Hardcastle, Birkinshaw \\& Worrall 1998a). The full complexity is illustrated by X-ray observations of the nearby powerful cluster radio galaxy Cygnus A. {\\it EXOSAT} and {\\it Ginga} have found evidence for highly obscured core emission, ($N_H \\sim 4 \\times 10^{23}$ cm$^{-2}$; Arnaud \\etal\\ 1987, Ueno \\etal\\ 1994) which should not have been seen with {\\it ROSAT}, given its low-energy X-ray passband. However, an unresolved core component was detected with the {\\it ROSAT} HRI (Harris, Perley \\& Carilli 1994), implying that Cygnus A's core has both an absorbed and an unabsorbed X-ray component; while the absorbed component may be associated with emission from the AGN, the unabsorbed component may be radio-related (Worrall 1997). Birkinshaw \\& Worrall (1993) argued in the case of NGC 6251 that a plausible source for such compact X-ray emission is synchrotron self-Compton (SSC) radiation from the base of the radio jet, originating on scales larger than that of the torus and so avoiding absorbtion. Such emission would be suppressed, but not eliminated, by relativistic beaming effects in radio galaxies and [as discussed in Worrall \\etal\\ (1994) and references therein] almost certainly dominates the X-ray emission in core-dominated quasars. It is therefore important in interpreting the X-ray emission from radio galaxies to have spectral or spatial information capable of distinguishing a non-thermal or compact component from thermal or extended emission. High-redshift powerful radio galaxies are important as counterparts to quasars in unified models, and Worrall \\etal\\ (1994) performed such a spatial separation for the $z = 1$ radio galaxy 3C\\,280. They showed that the unresolved component fell on an extension of the correlation between X-ray and radio core flux observed for core-dominated quasars, consistent with the model discussed above. In this paper we report X-ray observations of the radio galaxy 3C\\,220.1 with the {\\it ROSAT} HRI. We use the high resolution of the HRI to constrain the contributions from compact and extended emission. 3C\\,220.1 is an FRII (Fanaroff \\& Riley 1974) narrow emission-line radio galaxy at $z=0.61$ (Spinrad \\etal\\ 1985). With $H_0 = 50$ km s$^{-1}$ Mpc$^{-1}$ and $q_0 = 0$, used throughout the paper, its 178-MHz luminosity is $3.6 \\times 10^{27}$ W Hz$^{-1}$ sr$^{-1}$. Radio imaging (Burns \\etal\\ 1984; Jenkins, Pooley \\& Riley 1977; Harvanek \\& Hardcastle 1998) shows it to be a typical classical double object with largest angular size 35 arcsec (see Fig.\\ \\ref{contour}, inset); at this redshift one arcsecond corresponds to 8.94~kpc, so the projected linear size of 3C\\,220.1 is about 300 kpc. Burns \\etal\\ (1984) report an unusually prominent one-sided jet in the eastern lobe, and the radio core is also comparatively prominent, which may be an indication that the source is significantly affected by relativistic beaming, although no broad emission lines are reported by Spinrad \\etal\\ (1985). The prominent radio core, compared with the weak cores of the objects observed by Worrall \\etal\\ (1994), was the motivation for the present observations, since it allows us to probe the possible core X-ray and radio association. Optical observations show no evidence for a rich cluster near 3C\\,220.1, although the presence of a gravitational lens arc with $z=1.49$ in HST observations implies a deep potential well (M.\\ Dickinson, private communication, 1997). ", "conclusions": "We observed 3C\\,220.1 in order to test the hypothesis of Worrall \\etal\\ (1994) that all powerful radio sources exhibit a compact X-ray component related to their radio core. We find that there is strong evidence for a compact X-ray component in this object, that it is too bright to be attributed to a cooling flow under some simple assumptions, and that it follows the expected positive correlation between radio and X-ray core flux or luminosity. 3C\\,220.1 lies slightly above the line of slope unity plotted through the core-dominated quasars (Fig.\\ \\ref{cores}). Together with its unusually prominent radio core and one-sided radio jet, this may be an indication that we are viewing the source at an angle to the line of sight which is close to the radio galaxy-quasar boundary, so that its central X-ray emission includes a non-jet-related component originating close to the AGN. Extended X-ray emission is also unequivocally detected around 3C\\,220.1, with core radius and luminosity comparable to that of nearby rich clusters. This is qualitatively consistent with the HST detection of a gravitational lensing arc near the source." }, "9804/astro-ph9804334_arXiv.txt": { "abstract": "Deep infrared and optical images are presented of three proposed remnants of Thorne-$\\dot{\\rm Z}$ytkow Objects (T$\\dot{\\rm Z}$O). In particular, the IR data go several infrared magnitudes deeper than previous observations and in at least one case reveal the existence of weak objects within the error circles. It is argued, however, that none of the objects is likely to be the binary companion to the X-ray source in that region. These data present severe limits on any possible star or residual envelope at the distances of the respective X-ray objects. ", "introduction": "Recent work by several authors, especially van Paradijs et al. (1995) and Mereghetti \\& Stella (1995), have identified a group of pulsating X-ray sources that appear to have no detectable optical counterpart. In addition, these objects show no evidence of any binarity and all have pulse periods in the 5-10s range. Their X-ray luminosities are all sufficiently high that accretion must be occurring on to the neutron star from some source other than the interstellar medium. One possible scenario is that they are the first objects identifiable as the remains of Thorne-$\\dot{\\rm Z}$ytkow Objects (T$\\dot{\\rm Z}$Os). The progenitor T$\\dot{\\rm Z}$Os would have consisted of a neutron star embedded in the core of another star (Thorne \\&$\\dot{\\rm Z}$ytkow 1977). It is thought that these T$\\dot{\\rm Z}$Os could be formed as a result of the evolution of high mass systems through a common envelope phase (Ghosh et al. 1997). Once formed, thermonuclear burning via the {\\it r-p} process occurs in the core region of the T$\\dot{\\rm Z}$O. This supports the highly convective outer envelope against gravitational forces. However, once there is insufficient fuel remaining for the {\\it r-p} process to occur efficiently, the outer envelope collapses on the Kelvin-Helmholtz timescale (Podsiadlowski, Cannon and Rees, 1995). Some of the envelope material then forms into a disk around the neutron star with a radius that is dependent on the amount of angular momentum in the system. It is from this disk of material that the neutron star may accrete enough matter to power the X-ray pulsations seen. In order to explore this possibility deep infrared imaging observations have been carried out from UKIRT and the NASA IRTF to search for any IR emission from such a disk around three of the neutron stars. The observations reach as deep as J$\\sim$20 and in some cases reveal the presence of faint objects in the X-ray error circles. However, the characteristics of the IR colours, and other considerations, strongly suggest that these are not the TZO envelopes around the neutron stars. However, the new stringent limits on any IR companion in these systems leave these objects as enigmatic as ever. The three systems investigated here are: \\subsection{1E2259+586} The X-ray pulsar 1E2259+586 lies in the centre of the supernova remnat G109.1-1.0. The most recent estimates on its age make it approximately 3000 years old (Parmar {\\it et al.} 1997), though values up to 10000 years have been quoted. G109.1-1.0 is at an estimated distance of about 4 kpc. Fahlman{\\it et al.} (1982) demonstrated that the X-ray column densities to both the pulsar and the SNR were similar and hence that these two objects were probably related. However, the latest results from Parmar et al.(1997) indicate that this may not be the case. They find the X-ray column density to the pulsar to be greater than that to the supernova remnant. They suggest that this could be due to the pulsar being at a greater distance than the supernova remnant or to the presence of absorbing material local to the pulsar. However, assuming the same distance to 1E2259+586 as the SNR implies a pulsar X-ray luminosity of $\\approx10^{28}$ Watts. Results from {\\it EXOSAT}, {\\it GINGA} and {\\it BeppoSAX} (Hanson et al. 1988, Koyama et al. 1989 and Parmar et al. 1997 respectively) have shown that the 7s period of 1E2259+586 is slowly increasing, but the spindown is too slow to power the observed luminosity. There is no evidence of binary motion, the upper limit on $a_{x} \\sin i$ from pulse timing measurements being as low as 30 light-ms (Mereghetti, Israel and Stella, 1998). Early searches for the optical counterpart (Fahlman {\\it et al.} 1982; Margon \\& Anderson 1983) found several possible candidates, one of which (star D) was tentatively identified with 1E2259+586 by Middleditch, Pennypacker \\& Burns (1983) on the basis of IR pulsations. However, more sensitive fast photometric observations by Davies {\\it et al.} (1989) failed to find any pulsations, and there would now appear to be little positive evidence associating star D with 1E2259+586. A fainter, multicolour photometric study of all the possible counterparts by Davies \\& Coe (1991), together with revised astrometry, updated the list of possible counterparts. Coe \\& Jones (1992) presented the first optical spectroscopy of all the candidates brighter than V=23 and re-analysed the Einstein X-ray images of this source taken in 1981 in order to check the position and size of the X-ray error circle. They found that the relocation and reduction in the size of the error circle stemming from the re-processing of these data considerably changes the perspective on possible counterparts to the X-ray pulsar. In addition, a careful study of the images reveals previously unreported features that are possibly associated with jet activity from the pulsar, a possibility originally suggested by Gregory and Fahlman (1980). However, Hurford \\& Fesen 1995 report subsequent {\\it Rosat} HRI observations which, while confirming the existence of such features, they suggest are just statistical fluctuations in the SNR emission. \\subsection{4U0142+62} This persistent source has appeared in most X-ray catalogues and was originally confused (due to poor angular resolution) with the Rosat source RX J0146.9+6121 (=LSI +61 235) - the two sources are only 24 arcminutes apart. Measurements by White et al. (1987) with {\\it EXOSAT} reported a 25m modulation from the region (later shown to be coming from LSI +61 235 (Mereghetti, Stella \\& De Nile 1993)) but using the unmodulated lower energy data they were able to identify the position of 4U0142+62. The detection of X-ray pulsations at 8.7s was reported by Israel et al. (1994) who concluded that they were most likely coming from the optically unidentified system 4U0142+62 rather than the new {\\it Rosat} source. Further measurements by White et al. (1996) using {\\it ASCA} confirmed this conclusion and also demonstrated evidence for an excess in the X-ray halo around the source. This could be due to material local to the pulsar along the line of site, possibly in the form of a molecular cloud or perhaps the remains of a TZO envelope. No evidence of any binary motion has been reported from the X-ray signal. The accurate position reported by White et al. (1987) allowed them to locate the X-ray error circles from both {\\it EXOSAT} and {\\it Einstein} on the sky. As in the case of 1E2259+586 there is no evidence of any obvious optical candidates within the defined regions. They obtained an R band image and set a limit of R$\\geq$22.5 for any counterpart. The ASCA observations (White et al. 1996) determined the column to the source to be 8 x 10${^{21}}$ cm${^{-2}}$ which in turn allows them to deduce an optical extinction of A$_{v}$$\\sim$4.7. The distance to the source is not well defined but probably lies in the range 0.5--2.0 kpc. As in the case of 1E2259+586 a molecular cloud lies near to, or in front of 4U0142+62 (Mereghetti \\& Stella 1995) and may be affecting the column measurement. \\subsection{RX J1838.4--0301} This source was originally reported by Schwentker (1994) as 5.5s pulsar possibly associated with a SNR at a distance of 4kpc. Interestingly, in this case the X-ray positional error circle was reported to include a $\\sim$14 mag optical candidate. Subsequently Mereghetti et al. (1997) reanalysed the same {\\it Rosat} data and reported are different interpretation of the data. They obtained an optical spectrum of the star in the error circle and concluded that this object was simply a K5 main sequence star and that the X-ray emission originated from its corona . ", "conclusions": "The first deep infrared and optical images have been presented of three proposed remnants of Thorne-$\\dot{\\rm Z}$ytkow Objects (T$\\dot{\\rm Z}$O). In particular, the IR data go several infrared magnitudes deeper than previous observations and in at least one case, 1E2259+586, reveal the existence of weak objects within the X-ray error circle. It is argued, however, that in that case none of the objects seen is likely to be the binary companion to the X-ray source in that region. These data present new limits on any possible IR counterpart to the respective X-ray objects." }, "9804/astro-ph9804272_arXiv.txt": { "abstract": "Recently Alard\\markcite{alard1996} proposed to detect the shift of a star's image centroid, $\\delta x$, as a method to identify the lensed source among blended stars. Goldberg \\& Wo\\'zniak\\markcite{goldberg1997} actually applied this method to the OGLE-1 database and found that 7 out of 15 events showed significant centroid shifts of $\\delta x \\gtrsim 0.2$ arcsec. The amount of centroid shift has been estimated theoretically by Goldberg.\\markcite{goldberg1997} However, he treated the problem in general and did not apply it to a particular survey or field, and thus based his estimates on simple toy model luminosity functions (i.e., power laws). In this paper, we construct the expected distribution of $\\delta x$ for Galactic bulge events by using the precise stellar LF observed by Holtzman et al.\\markcite{holtzman1998} using HST. Their LF is complete up to $M_I\\sim 9.0$ ($M_V\\sim 12$), corresponding to faint M-type stars. In our analysis we find that regular blending cannot produce a large fraction of events with measurable centroid shifts. By contrast, a significant fraction of events would have measurable centroid shifts if they are affected by amplification-bias blending. Therefore, Goldberg \\& Wo\\'zniak's measurements of large centroid shifts for a large fraction of microlensing events confirms the prediction of Han and Alard that a large fraction of Galactic bulge events are affected by amplification-bias blending. ", "introduction": "Experiments to detect Massive Astronomical Compact Halo Objects (MACHOs) by monitoring the light variations of stars undergoing gravitational microlensing events are being carried out by several groups (MACHO, Alcock et al.\\ 1997a; EROS, Ansari et al.\\ 1996; OGLE, Udalski et al.\\ 1997; DUO, Alard \\& Guibert \\ 1997).\\markcite{alcock1997a, ansari1996, udalski1997, alard1997b} Since the lensing probability for a single source star is very low, these searches are being conducted toward very dense star fields such as the Large Magellanic Cloud and Galactic bulge. While searches towards these crowded fields result in an increased event rate, it also implies that many of the observed light curves include light from unresolved stars that are not being lensed: the blending problem. Depending on the source for the blended light, blending can be classified into several types. The first type, ``regular blending'', occurs when a bright source star registered on the template plate is lensed and its flux is blended with the light from numerous faint unresolved stars below the detection limit imposed by crowding. Regular blending affects the results of lensing experiments in various ways. First, it makes the measured event timescale shorter than the true one. Since the lens mass scales as $M \\propto t_{\\rm E}^{2}$, the lens mass determined from the measured timescales will be underestimated and the lens population will be misinterpreted. In addition, since the optical depth is directly proportional to the summation of event timescales, i.e., $\\tau \\propto \\sum_i t_{{\\rm E},i}$, the Galactic MACHO fraction determined from the optical depth without a proper blending correction is subject to great uncertainty (Di Stefano \\& Esin 1995).\\markcite{distefano1995} The second type of blending occurs if the source for the blended light is the lens itself: ``lens blending'' (Kamionkowski 1995; Buchalter, Kamionkowski, \\& Rich 1996; Buchalter \\& Kamionkowski 1997; Alard 1997).\\markcite{kamionkowski1995, bucjalter1996, buchalter1997, alrad1997} Besides the effects of regular blending, lens blending has several additional effects on the result of lensing experiments. First, because detecting events due to lenses close to the observer is comparatively more difficult than detecting events produced by lenses near the source, lens blending makes the optical depth depend on the geometry of the lens system. As a result, the matter distribution derived from the optical depth distribution deviates from its true one. Secondly, lens blending causes the lensing optical depth to depend on the lens mass function since more massive stars, which contribute more to the total optical depth, tend to be brighter, resulting in a larger blending effect (Nemiroff 1997; Han 1998).\\markcite{nemiroff1997, han1998} Finally, ``amplification bias'' blending occurs when one of several faint stars in the seeing disk below the detection limit is lensed, and the flux of the lensed star is associated with the flux from other stars in the integrated seeing disk (Nemiroff 1997).\\markcite{nemiroff997} In current experiments, photometry is carried out by comparing template images, in which only very bright stars are resolved and registered, with a series of images taken of the same field. The result is that in amplification-biased events the brightest star appears to be the source because the lensed star is too faint to be resolved. To be detected, the amplification-biased event must be highly amplified to overcome the high threshold flux from the brighter star. Therefore, the mean detection probability to detect events for each source star will be very low. However, if these faint stars comprise a significant fraction of the total number of stars, a considerable fraction of events might be amplification-biased (Han 1997b; Alard 1997a).\\markcite{han1997b, alard1997a} Moreover, due to the large amount of blended light from a bright star, the effects of blending for these events would be much more significant than those caused by other blending types. There have been several methods proposed to correct for the blending problem. The first method is to introduce an additional lensing parameter representing the residual flux from unresolved faint stars into the light curve fitting process. However, this method suffers from large uncertainties in the derived lensing parameters as a result of parameter degeneracies (Wo\\'zniak \\& Paczy\\'nski 1997).\\markcite{wozniak1997} Early-warning systems (for MACHO: Pratt et al.\\ 1996; for OGLE: Udalski et al.\\ 1994; for PLANET: Albrow et al.\\ 1995)\\markcite{pratt1996, udalski1994, albrow1995} allow one to construct lensing light curves with high time resolution and small photometric errors, enabling one to detect small color shifts caused by blending (Buchalter et al.\\ 1996)\\markcite{buchalter996}. However, due to the narrow distribution of colors for Galactic bulge stars, the expected color shifts are small. Han (1997b)\\markcite{han1997} proposed to use the Hubble Space Telescope (HST) to provide blending corrections. By using the high resolving power of HST combined with color information from ground-based observations, one can identify the lensed source star in the blended seeing disk, thus the uncertainty in the derived timescale can be significantly reduced. However, this method requires costly HST times. One can also correct the blending effect statistically if the luminosity function(LF) of stars well below the detection limit can be constructed (Alcock et al.\\ 1997b)\\markcite{alcock1996}, but in this case we lose information about individual events. Recently Alard (1996)\\markcite{alard1996} proposed to detect the shift of a star's image centroid, $\\delta x$, as a method to identify the lensed source among blended stars. Goldberg \\& Wo\\'zniak (1997)\\markcite{goldberg1997} actually applied this method to the OGLE-1 database and found that 7 out of 15 events showed significant centroid shifts of $\\delta x \\gtrsim 0.2$ arcsec. The amount of centroid shift has been estimated theoretically by Goldberg (1997).\\markcite{goldberg1997} However, he treated the problem in general and did not apply it to a particular survey or field, and thus based his estimates on simple toy model luminosity functions (i.e., power laws). In this paper, we construct the expected distribution of $\\delta x$ for Galactic bulge events by using the precise stellar LF observed by Holtzman et al.\\ (1998)\\markcite{holtzman1998} using HST. Their LF is complete up to $M_I\\sim 9.0$ ($M_V\\sim 12$), corresponding to faint M-type stars. In our analysis we find that regular blending cannot produce a large fraction of events with measurable centroid shifts. By contrast, a significant fraction of events would have measurable centroid shifts if they are affected by amplification-bias blending. Therefore, Goldberg \\& Wo\\'zniak's measurements of large centroid shifts for a large fraction of microlensing events confirms the prediction of Han (1997a) and Alard (1997) that a large fraction of Galactic bulge events are affected by amplification-bias blending. ", "conclusions": "The fact that events affected by amplification-bias blending produce large $\\delta x$ while the centroid shifts caused by regular blending are small can be understood analytically in the following way. To produce centroid shifts large enough to be measured, events should satisfy two conditions. First, the event should be highly amplified. For a very low amplification event ($A_{\\rm abs}\\sim 1$), the expected centroid shift will be small since $A_{\\rm abs}-1 \\sim 0$ and thus $\\eta \\sim 0$ in equation (1). On the other hand, for a very high-amplification event ($A_{\\rm abs}\\sim \\infty$), one finds $\\delta x\\sim \\left\\vert\\langle {\\bf x}\\rangle -{\\bf x}_{j}\\right\\vert$ since the factor $\\eta$ approaches unity. However, not all events that satisfy the first condition produce large centroid shifts. The second condition is that the lensed star should be one of the faint stars in the blended seeing disk. If the lensed star is the brightest one in the effective seeing disk and its flux dominates the flux over those from other faint blended stars, i.e., $\\kappa\\sim 1$, the position of the CL before gravitational amplification will be very close to that of the lensed star, resulting in a small amount of shift ($\\left\\vert\\langle {\\bf x}\\rangle-{\\bf x}_j\\right\\vert \\sim 0$) since $\\sum_{i\\neq j} F_{0,i}{\\bf x}_i + F_{0,j}{\\bf x}_j\\sim F_{0,j}{\\bf x}_j$ and $\\sum_i F_{0,i} \\sim F_{0,j}$. On the other hand, if the lensed source is very faint, i.e., $\\kappa\\sim 0$, the light from the source star has negligible effect on the position of CL, resulting in high possibility of a large centroid shift. In summary, the expected centroid shifts for various extreme cases of amplification and source star brightness are: $$ \\cases{ \\eta\\sim 0\\ {\\rm and}\\ \\delta x\\sim 0 & for a low amplification event \\cr & ($A_{\\rm abs}\\sim 1$) \\cr \\delta x\\sim \\left\\vert\\langle{\\bf x}\\rangle-{\\bf x}_j\\right\\vert\\sim 0 & for a high-amplification event\\cr & with luminous source \\cr & ($A_{\\rm abs}\\sim \\infty$, $\\kappa\\sim 1$) \\cr \\delta x\\sim \\left\\vert\\langle{\\bf x}\\rangle-{\\bf x}_j\\right\\vert & for a high-amplification event\\cr & with faint source \\cr & ($A_{\\rm abs}\\sim \\infty$, $\\kappa\\sim 0$). \\cr } \\eqno(4) $$ For amplification-biased events, source stars are in general very faint ($\\kappa\\sim 0$), mostly far below the detection limit. Despite their low luminosities, the fact that they are detected implies that the source stars are highly amplified ($A_{\\rm abs}\\sim \\infty$). Therefore, the conditions for large centroid shifts agree well with those for amplification-biased events. On the other hand, regular blended events do not meet these conditions. First, due to relatively small amount of blended light, regular blended events do not need to be highly amplified for detection ($A_{\\rm abs}\\sim 1$). Although they can be highly amplified ($A_{\\rm abs}\\sim \\infty$), the dominance of their fluxes ($\\kappa\\sim 1$) over those from other faint blended stars will result in small $\\delta x$. As demonstrated by the large centroid shifts for a significant fraction of events, the effect of amplification-bias blending on the results of lensing experiments is important. However, the methods to correct for the blending effect mentioned in \\S\\ 1 have various limitations in application. One very simple but very practical method to minimize the effect of amplification-bias blending is to monitor only very bright stars. With increasing reference image brightness, the required amplification for detection becomes higher, resulting in a lower probability of amplification-biased events. To show how the blending effect decreases with increasing reference image brightness, we simulate Galactic bulge events which are expected to be detected for various threshold reference image brightnesses. For each event, we compute the light fraction of the source star $\\kappa$ and the timescale decrease factor $\\eta$. The distributions of $\\kappa$ and $\\eta$ are shown in the upper panel of Figure 6. For a given fraction of source star flux, the observed timescale is reduced by $\\eta=t_{\\rm eff}/t_{\\rm E}= \\left[ 2(1-A_{\\rm min}^{-2})^{-1/2}-2 \\right]^{1/2}; \\ A_{\\rm min} = 0.34/\\kappa + 1$. Here events are assumed to be detected as long as they can amplify the reference image flux by more than a factor of 1.34. However, highly blended events will have short $t_{\\rm eff}$, resulting in low detectability. Therefore, we correct the distributions by the detection efficiency. We assume that the detection efficiency is linearly proportional to the timescale decrease factor, i.e., $\\epsilon \\propto \\eta$. The efficiency-corrected distributions $f(\\kappa)$ and $f(\\eta)$ are presented in the middle panels. In the lower panels we present the distributions of the fraction of events with $\\kappa \\ge \\kappa_{\\rm lim}$, $1-\\int_{0}^{\\kappa_{\\rm lim}} f(\\kappa)d\\kappa$, and $\\eta \\ge \\eta_{\\rm lim}$, $1-\\int_{0}^{\\eta} f(\\eta')d\\eta'$. From these distributions one finds that under the current threshold reference image brightness of $M_I\\sim 3$, the fraction of events with little blending effects ($\\kappa \\gtrsim 0.9$ or $\\eta \\gtrsim 0.9$) is $\\lesssim 10\\%$. However, as the brightness of the threshold reference image increases, this fraction gradually increases until it becomes $\\sim 80\\%$ when only stars brighter than $M_I\\sim 0$ are monitored, which corresponds to the brightness of Galactic bulge clump giant stars (Paczy\\'nski \\& Stanek 1998).\\markcite{paczynski1998} The MACHO group (1997a)\\markcite{alcock1997a} already applied this method and their optical depth determination is based on clump giant stars. However, by monitoring significantly fewer stars at a decreased event rate, the statistical precision of the lensing experiments will be lowered. A more general solution for the blending correction is provided by the rapidly progressing image subtraction technique (Alard \\& Lupton 1997b; Tomaney 1998)\\markcite{alard1997c, tomaney1998} which is also being applied to detect microlensing events towards M31 by the Colombia-Vatt group (Crotts \\& Tomaney 1996; Tomaney \\& Crotts 1996).\\markcite{crotts1996, tomaney1996}" }, "9804/astro-ph9804050_arXiv.txt": { "abstract": "We investigate if the gamma ray halo, for which recent evidence has been found in EGRET data, can be explained by neutralino annihilations in a clumpy halo. We find that the measured excess gamma ray flux can be explained through a moderate amount of clumping in the halo. Moreover, the required amount of clumping implies also a measureable excess of antiprotons at low energies, for which there is support from recent measurements by the BESS collaboration. The predicted antiproton fluxes resulting from neutralino annihilations in a clumpy halo are high enough to give an excess over cosmic-ray produced antiprotons also at moderately high energies (above a few GeV). This prediction, as well as that of one or two sharp gamma lines coming from annihilations into $\\gamma\\gamma$ or $Z\\gamma$ can be tested in upcoming space-borne experiments like AMS and GLAST. ", "introduction": " ", "conclusions": "" }, "9804/astro-ph9804099_arXiv.txt": { "abstract": "We investigate the nature of stellar populations of major galaxy mergers between late-type spirals considerably abundant in interstellar medium by performing numerical simulations designed to solve both the dynamical and chemical evolution in a self-consistent manner. We particularly consider that the star formation history of galaxy mergers is a crucial determinant for the nature of stellar populations of merger remnants, and therefore investigate how the difference in star formation history between galaxy mergers affects the chemical evolution of galaxy mergers. We found that the rapidity of star formation, which is defined as the ratio of the dynamical time-scale to the time-scale of gas consumption by star formation, is the most important determinant for a number of fundamental characteristics of stellar populations of merger remnants. Main results obtained in this study are the following five. (1) A galaxy merger with more rapid star formation becomes elliptical with larger mean metallicity. This is primarily because in the merger with more rapid star formation, a smaller amount of metal-enriched gas is tidally stripped away during merging and consequently a larger amount of the gas can be converted to stellar component. This result demonstrates that the origin of the color-magnitude relation of elliptical galaxies can be closely associated with the details of merging dynamics which depends on the rapidity of star formation in galaxy mergers. (2) Negative metallicity gradient fitted reasonably well by power-low can be reproduced by dissipative galaxy mergers with star formation. The magnitude of metallicity gradient is larger for an elliptical galaxy formed by galaxy merging with less rapid star formation. (3) Absolute magnitude of metallicity gradient correlates with that of age gradient in galaxy mergers in the sence that a merger remnant with steeper negative metallicity gradient is more likely to show steeper age gradient. (4) The outer part of stellar populations is both older and less metal-enriched than nuclei in an elliptical galaxy formed by galaxy merging with less rapid star formation. Moreover, the metallicity of the outer part of gaseous component for some models with less rapid star formation is appreciably smaller than that of stellar one. This result implies that the origin of metal-poor hot gaseous $X$-ray halo in real elliptical galaxies can be essentially ascribed to the dynamics of dissipative galaxy merging. (5) Irrespectively of the rapidity of star formation, the epoch of galaxy merging affects both the mean stellar metallicity and mean stellar age of merger remnants: Later galaxy mergers are more likely to become ellipticals with both younger and more metal-enriched stellar populations. This result reflects the fact that in the later mergers, a larger amount of more metal-enriched interstellar gas is preferentially converted into younger stars in the later star formation triggered by galaxy merging. These five results clearly demonstrate that even the chemical evolution of elliptical galaxies can be strongly affected by the details of dynamical evolution of galaxy merging, which is furthermore determined by the rapidity of star formation of galaxy mergers. In particular, tidal stripping of interstellar gas and total amount of gaseous dissipation during galaxy merging are demonstrated to play a vital role in determining a number of chemical properties of merger remnants. Based upon these results, we adopt a specific assumption of the luminosity dependence of the rapidity of star formation and thereby discuss how successfully the present merger model can reproduce a number of fundamental chemical, photometric, and spectroscopic characteristics of elliptical galaxies. ", "introduction": "Elliptical galaxies have been generally considered to be old, coeval and homogeneous systems passively evolving after the single initial burst of star formation associated with dissipative galaxy formation. This classical picture of elliptical galaxy formation appears to have been supported by the considerably tight color-magnitude relation of elliptical galaxies ( Bower, Lucey, \\& Ellis 1992; Ellis et al. 1997) and by relatively smaller redshift evolution of photometric properties of elliptical galaxies (Arag$\\rm \\acute{o}$n-Salamanca et al. 1993; Franx \\& van Dokkum 1996). A growing number of recent observational results, however, shed a strong doubt on this long-standing view of elliptical galaxy formation, and suggest that there is great variety of star formation history between elliptical galaxies, such as the epoch of major star formation, the duration and efficiency of star formation (Worthey, Faber, \\& Gonzalez 1992; Matteuchi 1994; Faber et al. 1995; Bender 1996; Worthey, Trager, \\& Faber 1996). This tendency that elliptical galaxies show diversity in star formation history and nevertheless can actually keep the tightness of the color-magnitude relation is considered to be quite mysterious and thus to provide any theoretical models with a valuable insight on the elliptical galaxy formation. Such kind of mysterious nature observed in elliptical galaxies is demonstrated to hold equally for the dynamical and kinematical properties of elliptical galaxies. For example, considerably small thickness of the fundamental plane of elliptical galaxies implies a rather smaller range of admitted dynamical state of the galaxies (Djorgovski \\& Davis 1987; Dressler et al. 1987 ; Djorgovski, Pahre, \\& de Carvalho 1996) whereas the morphological dichotomy between boxy-disky elliptical galaxies (Kormendy \\& Bender 1996) and the projected density profile systematically departing from de Vaucouleurs $R^{1/4}$ law (Caon, Capaccioli, \\& D'Onofrio 1993) show a great variety of major orbit families consisting the galaxies. These fundamental characteristics that elliptical galaxies show both diversity and uniformity in their chemical, photometric and dynamical properties have imposed some stringent but valuable constraints on any theoretical models of elliptical galaxy formation. What is the most vital in challenging the origin of elliptical galaxy formation in this kind of situation is to investigate whether or not both the chemical and photometric properties and dynamical and kinematical ones can be reproduced successfully by a specific model of galaxy formation in a reasonably self-consistent manner. The previous theoretical models addressing this important issue on elliptical galaxy formation are divided basically into two categories: The dissipative galactic collapse model (e.g., Larson 1976; Carlberg 1984) and the galaxy merger model (e.g., Toomre \\& Toomre 1972). As is suggested by Kormendy \\& Sanders (1992), these two dominant and apparently competing scenarios for elliptical galaxy formation are now converging, thus it would be crucial to construct one more realistic and sophisticated model of elliptical galaxy formation. Although there are a large number of important studies exploring the origin of elliptical galaxy formation along the dissipative collapse scenario, especially in the context of the nature of stellar populations (e.g., Arimoto \\& Yoshii 1987), we here restrict ourselves to the merger scenario of elliptical galaxy formation. Recent extensive studies of merger models of elliptical galaxy formation, mostly based upon numerical simulations, $appear$ to have succeeded in resolving most of the outstanding problems related to dynamical and kinematical properties of elliptical galaxies, such as the phase space density (Ostriker 1980; Carlberg 1986) and kinematical misalignment (Franx, Illingworth, \\& de Zeeuw 1991; Barnes 1992), by invoking the inclusion of bulge component, gaseous dissipation, and multiplicity of galaxy merging (Hernquist, Spergel, \\& Heyl 1993; Weil \\& Hernquist 1996; Barnes \\& Hernquist 1996). Although it would be safe to say that the galaxy merging between two late-type spirals is one of the most promising candidates explaining more clearly the origin of elliptical galaxies at least in the context of the $dynamical$ $and$ $kinematical$ $properties$, however, there still remain a number of unresolved and apparently serious problems concerning the merger model (e.g., van den Bergh 1995). One of the most crucial problems among these is on whether the fundamental characteristics of stellar populations of elliptical galaxies can be reproduced reasonably well by galaxy merging between two late-types spirals. Surprisingly, there are only a few works addressing this critical issue for the merger model, probably because it is considered to be rather difficult to solve the chemical evolution of galaxy mergers in which a number of competing physical processes are expected to affect strongly the chemical evolution of galaxy mergers. White (1980) and Mihos \\& Hernquist (1994) found that the stellar populations of progenitor disks are not mixed so well even by the violent relaxation during galaxy merging and consequently the metallicity gradient of progenitor disks is not so drastically washed out. The metallicity gradient of merger remnant is furthermore found not to be fitted by power law observed in elliptical galaxies (Mihos \\& Hernquist 1994). Schweizer \\& Seitzer (1992) discussed whether or not the bluer integrated $UBV$ color of elliptical galaxies with morphologically fine structure can be explained by secondary starburst induced by major disk-disk galaxy mergers. Kauffmann \\& Charlot (1997) construct a semi-analytic model of elliptical galaxy formation, which is based upon the hierarchical clustering in CDM universe and includes rather simple chemical enrichment process, and thereby demonstrate that the origin of the color-magnitude relation of elliptical galaxies can be reproduced successfully even in the CDM model of galaxy formation (See also Baugh, Cole \\& Frenk 1996.). Thus, since there are only a few works addressing chemical and photometric properties for the merger model, it is essential for the merger model to investigate more throughly the fundamental chemical and photometric properties of merger remnants, including the origin of color-magnitude relation (Faber 1973; Visvanathan \\& Sandage 1977), age and metallicity gradient (Peletier et al. 1990; Davies et al 1991), $\\rm Mg_{2} - \\sigma$ relation (Burstein et al. 1988), age-metal-conspiracy in stellar populations (Faber et al. 1995; Worthey et al. 1996), luminosity dependence of the line ratio [Mg/Fe] (Worthey et al. 1992), metal-poor gaseous $X$-ray halo (Matsumoto et al. 1997), and the substantially metal-enriched galactic nuclei at higher redshift (Hamann \\& Ferland 1993). What should be recognized foremost in investigating the nature of stellar populations in merger remnants is that a glowing number of observational results have been accumulated which suggest the relatively earlier formation of elliptical galaxies. Tightness of the color-magnitude relation in the cluster of galaxies (Bower et al. 1992, Ellis et al 1996), relatively smaller photometric evolution of cluster ellipticals (Arag$\\rm \\acute{o}$n-Salamanca et al. 1993), and the redshift evolution of the fundamental plane (Franx \\& van Dokkum 1996) all suggest the $typical$ formation epoch of elliptical galaxies is earlier than 2 in redshift. Furthermore, as is suggested by Kormendy \\& Sanders (1992), the fact that no galaxy in the $K$-band survey of Cowie et al. (1994) shows the global color resembling that of the Arp 220, which is considered to be ongoing mergers and forming ellipticals, implies that the formation epoch of elliptical galaxies should be earlier than 1.0 in redshift. Silva \\& Bothun (1997) revealed that the fraction of mass of stellar populations with intermediate age to total mass in elliptical galaxies with morphologically fine structure is less than 15 percent. These results imply that if elliptical galaxies are formed by galaxy merging, the epoch of galaxy merging should be relatively earlier and furthermore that the precursor disks of galaxy mergers may be extremely abundant in interstellar medium compared with the present spirals. Recent high quality imaging using $Hubble$ $Space$ $Telescope$ ($HST$) has revealed that a larger number of galaxies at faint magnitude are interacting/merging galaxies (e.g., van den Bergh et al. 1996), indicating furthermore that the potential candidate for elliptical galaxies formed by galaxy merging are ubiquitous in higher redshift universe. Hence it is quite reasonable and essential to study the nature of stellar populations of higher redshift galaxy mergers between disk galaxies with the gas mass fraction larger than 0.2, which is a typical value of the present late-type spirals, and thereby to confirm whether or not elliptical galaxies can be formed $actually$ by galaxy merging. The purpose of this paper is to explore the nature of the stellar populations of a gas-rich disk merger which is considered to be occurred the most frequently in the high redshift universe. We particularly investigate how successfully galaxy mergers between gas-rich spirals can reproduce a number of fundamental chemical, photometric, and spectroscopic properties of elliptical galaxies. The layout of this paper is as follows. In \\S 2, we summarize numerical models used in the present study and describe in detail methods for analyzing the stellar populations produced by dissipative galaxy mergers with star formation. In \\S 3, we demonstrate how a number of fundamental characteristics of stellar populations in merger remnants are affected by the star formation history of dissipative galaxy merging. In \\S 4, we discuss how successfully the present merger model can reproduce a number of observational results concerning the chemical, photometric, and spectroscopic properties of elliptical galaxies. In this section, we also point out the advantages and disadvantages of galaxy mergers in explaining both the chemical, photometric, and spectroscopic properties and dynamical and kinematical ones in real elliptical galaxies. The conclusions of the present study are given in \\S 5. ", "conclusions": "Main results obtained in the present study are summarized as follows. (1) Galaxy mergers with more rapid star formation become ellipticals with larger mean stellar metallicity, primarily because in the mergers with more rapid gas consumption, a smaller amount of metal-enriched gas is tidally stripped away during merging and consequently a larger amount of the gas can be converted into stellar component. This result is demonstrated not to depend so strongly on the other parameters such as the orbit configuration of galaxy merging and multiplicity of the mergers. These results suggest that the origin of the color-magnitude relation of elliptical galaxies can be closely associated with the details of merging dynamics which depends on the rapidity of star formation (thus on the galactic luminosity) in galaxy mergers. (2) Negative metallicity gradient fitted reasonably well by power-low can be reproduced by dissipative galaxy mergers with star formation, which is in good agreement with the recent observational results of elliptical galaxies. The absolute magnitude of metallicity gradient in each merger remnant depends on the orbit configuration of each galaxy merging, suggesting that the observed dispersion in the absolute magnitude of metallicity gradient for a given luminosity range of elliptical galaxies reflects the diversity in the orbit configuration of galaxy merging. (3) Absolute magnitude of metallicity gradient correlates with that of age gradient in a merger in the sence that a merger remnant with steeper negative metallicity gradient is more likely to show steeper age gradient. This result reflects the fact that the degree of violent relaxation and gaseous dissipation during merging strongly affect both the age gradient and metallicity one. (4) The outer part of stellar populations is both older and less metal-enriched than nuclei in an elliptical galaxy formed by galaxy merging with less rapid star formation. Moreover galaxy mergers with less rapid star formation are more likely to become ellipticals with metal-poor gaseous halo. This result suggests that the formation of metal-poor $X$-ray halo actually observed in elliptical galaxies can be essentially ascribed to the dissipative galaxy merging between late-type spirals, and furthermore provides a clue to a solution for the iron abundance discrepancy problem in elliptical galaxies. (5) The epoch of galaxy merging affects both the mean stellar metallicity and the mean stellar age in merger remnants: Later galaxy mergers become ellipticals with both younger and more metal-enriched stellar populations. This result suggests that the origin of Worthey's 3/2 rule (Worthey et al. 1996), which is invoked in maintaining the tightness of the color-magnitude relation of elliptical galaxies, can be understood in terms of the difference in the epoch of galaxy formation and transformation, that is, the epoch of galaxy merging, between elliptical galaxies. (6) Luminosity dependence of chemical, photometric, and spectroscopic properties in merger remnants, which is derived by adopting a specific assumption on the luminosity dependence of the rapidity of star formation of galaxy mergers, does not match so reasonable well with that observed in real elliptical galaxies. This result implies that other fundamental physical processes expected to be dependent on the galactic luminosity should be incorporated into the present merger model for more successful comparison with observational trends of luminosity-dependent chemical, photometric, and spectroscopic properties of elliptical galaxies. (7) As is described in the above (1) - (6), the details of gas dynamics of galaxy merging, in particular, the tidal stripping of metal-enriched interstellar gas and the degree of gaseous dissipation during merging, both of which depend on the star formation history of galaxy mergers, are demonstrated to determine even the chemical and photometric properties of merger remnants. These results can not be obtained until both the chemical and dynamical evolution during galaxy merging are solved numerically in a reasonably self-consistent way." }, "9804/astro-ph9804320_arXiv.txt": { "abstract": "Physical properties of the atomic gas in spiral galaxies are briefly considered. Although both Warm (WNM, 10$^4$~K) and Cool (CNM, $\\sim$~100 K) atomic phases coexist in many environments, the dominant mass contribution within a galaxy's star-forming disk (R$_{25}$) is that of the CNM. Mass fractions of 60 to 90\\% are found within R$_{25}$. The CNM is concentrated within moderately opaque filaments with a face-on surface covering factor of about 15\\%. The kinetic temperature of the CNM increases systematically with galactocentric radius, from some 50 to 200~K, as expected for a radially declining thermal pressure in the galaxy mid-plane. Galaxies of different Hubble type form a nested distribution in T$_K$(R), apparently due to the mean differences in pressure which result from the different stellar and gas surface densities. The opaque CNM disappears abruptly near R$_{25}$, where the low thermal pressure can no longer support the condensed atomic phase. The CNM is apparently a prerequisite for star formation. Although difficult to prove, all indications are that at least the outer disk and possibly the inter-arm atomic gas are in the form of WNM, which accounts for about 50\\% of the global total. Median line profiles of the CNM display an extremely narrow line core (FWHM $\\sim$~6~km~s$^{-1}$) together with broad Lorentzian wings (FWHM $\\sim$~30~km~s$^{-1}$). The line core is consistent with only opacity broadening of a thermal profile. The spatial distribution of CNM linewidths is not random, but instead is extremely rich in structure. High linewidths occur in distinct shell-like structures with diameter of 100's of pc to kpc's, which show some correlation with diffuse H$\\alpha$ shells. The primary source of ``turbulent'' linewidth in the atomic ISM appears to be organized motions due to localized energy injection on a scale of a few 100 pc. ", "introduction": "Over the years a succession of eminent authors has considered the question of which physical processes determine the temperature of neutral atomic gas and which timescales are required to achieve (a local) thermodynamic equilibrium. This began with Field, Goldsmith and Habing (1969) and continued with Draine (1978), Shull and Woods (1985) and most recently Wolfire {\\it et al.} (1995). While some of the processes involved are quite straightforward, others, like the photoelectric emission from dust grains, have had to be substantially updated to reflect our growing knowledge of the properties and abundance of interstellar dust. The consensus which has emerged is that heating is in fact dominated by the photoelectric heating from small dust grains over a wide range of conditions and environments. Cooling, on the other hand, is regulated primarily by emission in the [CII] 158~$\\mu$m fine structure line at densities in excess of about 1~cm$^{-3}$ and by Ly$\\alpha$ emission at lower densities. The relative importance of the various mechanisms is nicely illustrated in Fig.~3 of Wolfire {\\it et al.} (1995). In the same figure can be seen the characteristic phase diagram for atomic gas. Two thermodynamically stable phases are found. The first is the so-called Warm Neutral Medium (WNM) which predominates at low densities and pressures, and has a kinetic temperature that rises toward lower pressures from a value of perhaps 5000~K to 10$^4$~K, with an accompanying increase in ionization fraction, $x$~$\\sim$~0.1 to 0.9. The second is the Cool Neutral Medium (CNM) which is primarily found at high densities and pressures, with a kinetic temperature that decreases towards higher pressures from a value of some 200~K to perhaps as low as 20~K. Over some range of intermediate pressures thought to be typical of the interstellar medium in the local neighbourhood of the Galaxy (P/k$\\sim$ 2000~K~cm$^{-3}$), the two phases can coexist in pressure equilibrium. However, given the strong gradient in the mid-plane thermal pressure which is likely to follow from the combination of a radial exponential stellar disk and a similarly declining gas surface density distribution, we can safely predict that the inner regions of disk galaxies will have predominantly rather cool CNM, while the outer regions of galaxies will eventually be dominated by the WNM. At intermediate radii we might also expect to see a radial increase in the CNM kinetic temperature in response to the declining mid-plane pressure. ", "conclusions": "" }, "9804/astro-ph9804116_arXiv.txt": { "abstract": "Multidimensional cosmologies allow for variations of fundamental physical constants over the course of cosmological evolution, and different versions of the theories predict different time dependences. In particular, such variations could manifest themselves as changes of the proton-to-electron mass ratio $\\mu=m_{\\rm p}/m_{\\rm e}$ over the period of $\\sim 10^{10}$ yr since the moment of formation of high-redshift QSO spectra. Here we analyze a new, high-resolution spectrum of the $z = 2.81080$ molecular hydrogen absorption system toward the QSO PKS 0528--250 to derive a new observational constraint to the time-averaged variation rate of the proton-to-electron mass ratio. We find $| \\dot{\\mu}/ \\mu | < 1.5 \\times 10^{-14}$ yr$^{-1}$, which is much tighter than previously measured limits. ", "introduction": "The possibility of the variability of fundamental physical constants was first put forward by Dirac (1937) in the course of his discussion with Milne (1937). Later it was considered by Teller (1948), Gamow (1967), Dyson (1972) and other physicists. Interest in the problem increased greatly during the last decade, due to new developments in the Kaluza--Klein and supergravity models of unification of all the physical interactions. Chodos \\& Detweiler (1980), Freund (1982), Marciano (1984), and Maeda (1988) discussed possibilities of including these multidimensional theories into the cosmological scenario of the expanding Universe and found that the low-energy limits to the fundamental constants might vary over the cosmological time. Variations of the coupling constants of strong and electroweak interactions might then cause the masses of elementary particles to change. Note that an increase of the proton mass by 0.08\\% would lead to transformation of protons into neutrons (by electron capture), resulting in destruction of atoms in the Universe. As demonstrated by Kolb, Perry, \\& Walker (1987) and Barrow (1987), observational bounds on the time evolution of extra spatial dimensions in the Kaluza--Klein and superstring theories can be obtained from limits on possible variations of the coupling constants. Damour \\& Polyakov (1994) have developed a modern version of the string theory which assumes cosmological variations of the coupling constants and hadron-to-electron mass ratios. Therefore the parameters of the theory can be restricted by testing cosmological changes of these ratios. The present value of the proton-to-electron mass ratio is $\\mu=1836.1526645\\,(57)$ (CODATA, 1997). Obviously, any significant variation of this parameter over a small time interval is excluded, but such variation over the cosmological time $\\sim 1.5\\times 10^{10}$ yr remains a possibility. This possibility can be checked by analyzing spectra of high-redshift QSOs. The first analysis of this kind has been performed by Pagel (1977), who obtained a restriction $|\\dot{\\mu}/\\mu|< 5\\times 10^{-11}{\\rm ~yr}^{-1}$ on the variation rate of $\\mu$ by comparison of wavelengths of H\\,I and heavy-ion absorption lines, as proposed by Thompson (1975). This technique, however, could not provide a fully conclusive result, since the heavy elements and hydrogen usually belong to different interstellar clouds, moving with different radial velocities. In this paper we use another technique, based on an analysis of H$_2$ absorption lines only. One object suitable for such analysis is the $z = 2.811$ absorption system toward PKS 0528--250, in which Levshakov \\& Varshalovich (1985) identified molecular hydrogen absorption lines based on a spectrum obtained by Morton et al.\\ (1980). Foltz, Chaffee, \\& Black (1988) have presented a limit to possible variation of $\\mu$ based on their observations of PKS 0528--250. Their analysis did not, however, take into account wavelength-to-mass sensitivity coefficients, hence their result appeared to be not well grounded. Subsequently the spectrum of Foltz, Chaffee, \\& Black (1988) was reappraised by Varshalovich \\& Levshakov (1993), who obtained $|\\Delta\\mu/\\mu| < 0.005$ at the redshift $z=2.811$, and by Varshalovich \\& Potekhin (1995), who obtained $|\\Delta\\mu/\\mu| < 0.002$ at the $2\\sigma$ significance level. (Here $\\Delta\\mu/\\mu$ is the fractional variation of $\\mu$.) More recently, Cowie \\& Songaila (1995) used a new spectrum of PKS 0528$-$250 obtained with the Keck telescope to arrive at the 95\\% confidence interval $-5.5 \\times 10^{-4} < \\Delta\\mu/\\mu < 7 \\times10^{-4}$. Here we present a profile fitting analysis of a new, high-resolution spectrum of PKS 0528$-$250, obtained in November 1991 with the Cerro-Tololo Inter-American Observatory (CTIO) 4 m telescope. We have calculated the wavelength-to-mass sensitivity coefficients for a larger number of spectral lines and employed them in the analysis, which yields the strongest observational constraint yet to possible $\\mu$ variation over the cosmological time scale (eq.~[\\ref{limits}] below). ", "conclusions": "We have obtained a constraint to the variation rate of the proton-to-electron mass ratio $\\mu$. Two fitting procedures have been used, one of which simultaneously takes into account all observed spectral regions and transitions, while the other is applied to each spectral feature separately. The two techniques, applied to two different sets of spectral intervals, have resulted in similar upper bounds on $\\Delta\\mu/\\mu$, at the level $\\,\\sim2\\times10^{-4}$. The obtained restriction on $\\dot{\\mu}/\\mu$ (\\ref{constraint}) is by an order of magnitude more stringent than the limit set previously by Varshalovich \\& Potekhin (1995), who used a spectrum with a lower spectral resolution. Moreover, it is much more restrictive than the estimate of Cowie \\& Songaila (1995), based on high-resolution Keck telescope observations. There are two reasons for the higher accuracy of the present estimate. First, our fitting procedure simultaneously takes into account all observed spectral regions and transitions. This is particularly important because many of the transitions are blended, even at the spectral resolution of the spectrum used by Cowie \\& Songaila (1995). A separate analysis of spectral lines leads to larger statistical errors, as we have shown explicitly in Section 4.2. Second, we include a larger number of transitions between excited states of the H$_2$ molecule (83 spectral lines, compared with 19 lines used by Cowie \\& Songaila), many of which have higher wavelength-to-mass sensitivity coefficients $K_i$. The larger interval of $K_i$ values results in a higher sensitivity to possible mass ratio deviations. The method used here to determine the variation rate of $\\mu$ could be formally less sensitive than the one based on an analysis of relative abundances of chemical elements produced in the primordial nucleosynthesis (Kolb et al., 1986). However, the latter method is very indirect since it depends on a physical model which includes a number of additional assumptions. Therefore the present method seems to be more reliable. Quite recently, Wiklind \\& Combes (1997) used a similar method (following Varshalovich \\& Potekhin, 1996) in order to infer limits on time variability of masses of molecules CO, HCN, HNC and the molecular ion HCO$^+$ from high-resolution radio observations of rotational lines in spectra of a few low-redshift ($z<1$) quasars. The result reported in this paper constrains the mass of the H$_2$ molecule, and thus the proton mass, at much larger $z$. These constraints may be used for checking the multidimensional cosmological models which predict time-dependences of fundamental physical constants. The described method of the calculation of the sensitivity coefficients can also be used for analyzing any other high-redshift molecular clouds, which may be found in future observations." }, "9804/hep-ph9804291_arXiv.txt": { "abstract": "Mixed dark matter scenario can reconcile the COBE data and the observed large scale structure. So far the massive neutrino with a mass of a few eV has been the only discussed candidate for the hot dark matter component. We point out that the hadronic axion in the so-called hadronic axion window, $f_{a} \\sim 10^{6}$~GeV, is a perfect candidate as hot dark matter within the mixed dark matter scenario. The current limits on the hadronic axion are summarized. The most promising methods to verify the hadronic axion in this window are the resonant absorption of almost-monochromatic solar axions from M1 transition of the thermally excited $^{57}$Fe in the Sun, and the observation of the ``axion burst'' in water \\v{C}erenkov detectors from another supernova. ", "introduction": "The cold dark matter (CDM) dominated universe with scale-invariant primordial density fluctuation has been the standard theory of structure formation. After COBE has found the finite density fluctuation in the cosmic microwave background radiation (CMBR), the standard CDM scenario was found to give too much power on smaller scales. Many modifications to the standard CDM scenario were proposed which solve the discrepancy: by introducing a small Hot Dark Matter (HDM) component~\\cite{aph9707285}, by ``tilting'' the primordial density fluctuation spectrum~\\cite{tilt}, by assuming a finite cosmological constant~\\cite{Lambda-CDM}, or by introducing particles (such as $\\nu_{\\tau}$) whose decay changes the time of radiation-matter equality~\\cite{cdm+mnu}. At this point, there is no clear winner among these possibilities.\\footnote{However, a large ``tilt'' is difficult to obtain in many inflationary models. $\\tau$CDM can be tested well by $B$-factory experiments in the near future~\\cite{hph9709411}. The recent data from high-redshift supernovae prefer $\\Lambda$CDM \\cite{Saul}, but the possible evolution of supernovae needs to be excluded by more systematic comparison between nearby and high-$z$ supernovae.} In this letter, we revisit the mixed dark matter (MDM) scenario from the particle physics point of view. This scenario has attracted strong interests because there has been a natural candidate for the HDM component: massive neutrino(s). A neutrino with a mass of a few eV can naturally contribute to a significant fraction of the current universe. However, it has not been easy to incorporate the HDM together with other neutrino ``anomalies,'' unless all three generation neutrinos (possibly together with a sterile neutrino) are almost degenerate, and their small mass splittings explain various ``anomalies.'' Such a scenario may be viewed as fine-tuned. Especially, the atmospheric neutrino anomaly is quite significant statistically now thanks to the SuperKamiokande experiment, which suggests the mass squared difference of $\\Delta m^{2} = 10^{-3} - 10^{-2}~{\\rm eV}^{2}$ between the muon and tau neutrinos. If we view the situation from the familiar hierarchical fermion mass matrices, it suggests the tau neutrino mass of 0.03 -- 0.1~eV, and it appears difficult to accommodate the HDM based on massive neutrinos. We point out that the hadronic axion~\\cite{KSVZ} can be an alternative motivated candidate for the HDM component in the MDM model. Axion has been proposed as a solution to the strong CP problem in the QCD, and the hadronic axion (or KSVZ axion) is one version which predicts small coupling of the axion to the electron. There has been known a window of $f_{a} \\sim 10^{6}$~GeV allowed by existent astrophysical and cosmological constraints if the axion coupling to photons is suppressed accidentally. This is referred to as the ``hadronic axion window.'' Our main observation is that this window gives exactly the right mass of $m_{a} \\sim \\mbox{a few eV}$ and the number density of the axion appropriate for the HDM component in the MDM scenario. ", "conclusions": "In this letter, we have pointed out that the hadronic axion in the hadronic axion window ($f_a\\sim 10^6{\\rm ~GeV}$) can automatically be a good candidate of the Hot Dark Matter component in the mixed dark matter scenario. In order to evade an astrophysical constraint from the background UV light, axion-photon-photon coupling has to be suppressed in the hadronic axion window, probably by an accidental cancellation. This scenario may be tested by detecting the axion burst from a future supernova in water \\v{C}erenkov detectors, or detecting solar axions using resonant absorption." }, "9804/astro-ph9804071_arXiv.txt": { "abstract": "We present CCD BVI photometry of the old open cluster Berkeley 21, one of the most distant clusters in the Galactic anticentre direction, and possibly the lowest metallicity object in the open clusters sample. Its position and metal abundance make it very important for the study of the Galactic disc. Using the synthetic Colour - Magnitude Diagram method, we estimate values for distance modulus \\mmm = 13.4--13.6, reddening \\ebv = 0.74--0.78 (with possible differential absorption), and age = 2.2--2.5 Gyr. ", "introduction": "Old open clusters cover a large range of distances, metallicities, and ages (Friel 1995), and that warrants their use in investigations of the chemical and dynamical evolution of our Galaxy. To study the metallicity and age distribution of open clusters with Galactocentric distance, and avoid unnecessary and dangerous biases, a key requisite is homogenous analysis of very accurate observational data, as discussed by, e.g, Janes \\& Phelps (1994, JP94) Carraro \\& Chiosi (1994, CC94), Friel (1995), Twarog et al. (1997,TAAT97). This is the fifth paper of a series dedicated to the examination of old open clusters of different ages and metallicities, and located at different Galactic radii: for them we measure in a homogenous way distance, age, reddening and metallicity. These quantities are derived from comparison of the observed colour-magnitude diagrams (CMDs) to synthetic ones generated by a numerical code based on stellar evolution tracks and taking into account theoretical and observational uncertainties (Tosi et al. 1991). These simulations are much more powerful than the classical isochrone fitting method to study the evolutionary status of the analysed region and have been successfully applied both to nearby irregular galaxies (Greggio et al. 1998 and references therein) and to galactic open clusters (NGC2243: Bonifazi et al. 1990; Cr261: Gozzoli et al. 1996; NGC6253: Bragaglia et al. 1997; NGC2506: Marconi et al. 1997). Berkeley 21 (Be21) is located toward the Galactic anticentre, at coordinates RA(1950) = 5:48:42, DEC(1950) = 21:46, and l$_{\\rm II}$ =187$^{\\circ}$, b$_{\\rm II}$ = $-2.5^{\\circ}$. It has already been observed by Christian \\& Janes (1979, hereafter CJ), but their photographic CMD is very shallow, barely reaching the main sequence Turn-Off (TO). They deduced a substantial reddening (\\ebv $\\simeq$ 1.0), a large distance modulus (\\mmm $\\simeq$ 16), and a quite young age ($\\sim 10^8$ yr). Much better data have been presented by Phelps et al. (1994, PJM94) in their compilation of old open clusters, providing \\mmm = 13.9$\\pm$0.2 and an age of 2.8 Gyr, derived on the basis of $\\delta V$=1.6 ($\\delta V$ being the magnitude difference between TO and clump stars, JP94). The metallicity has been estimated by medium-resolution spectroscopy (Friel \\& Janes 1993, FJ93), but its actual value strongly depends on the adopted reddening (\\ebv = 0.7$\\pm$0.2, Janes 1991), with [Fe/H]= $-0.97^{+0.3}_{-0.1}$ dex. This large uncertainty, given the fact that Be21 defines the lowest metallicity limit of the open clusters sample and is one of the clusters most distant from the Galactic centre, is a further limitation for studies of the (possible) age and distance relations with chemical abundance in the Galactic disc (see also Twarog et al. 1997). In Section 2 we describe the observations and data analysis; in Section 3 we present the derived CMDs involving BVI photometry and discuss the presence of binary stars. In Section 4 we compare observed and synthetic CMDs and derive metallicity, age, distance and reddening. Finally, conclusions will be reviewed in Section 5. \\begin{table} \\begin{center} \\caption{Log of the observations. The cluster field has its centre at RA(2000) = 5:51:46, DEC(2000) = +21:48:45. The off-cluster field has coordinates: RA(2000) = 5:52:08, DEC(2000) = +21:53:53} \\begin{tabular}{lcccccccc} \\hline\\hline \\multicolumn{1}{c}{Night} &\\multicolumn{1}{c}{Field} &\\multicolumn{1}{c}{Tel.} &\\multicolumn{3}{c}{Exposure in seconds} &\\multicolumn{1}{c}{Seeing} \\\\ & & & B & V & I & excursion \\\\ \\hline Mar 4 &Centre &Danish & 120 & 120 & 120 & 0.90-1.20\\\\ Mar 4 &Centre &Danish & - & 120 & 900 & 0.85-1.00\\\\ Mar 4 &Centre &Danish & - & 600 & - & 0.95-1.10\\\\ Mar 5 &Centre &Danish & 120 & 120 & 120 & 1.00-1.10\\\\ Mar 5 &Centre &Danish &1500 & 900 & - & 0.95-1.10\\\\ Mar 14 &Ext. &Dutch & 120 & 60 & 60 & 1.10-1.30\\\\ Mar 14 &Ext. &Dutch & 1200 & 480 & 480 & 1.20-1.30\\\\ Mar 14 &Ext. &Dutch & 1200 & 480 & 480 & 1.15-1.30\\\\ \\hline \\end{tabular} \\end{center} \\label{tab-log} \\end{table} \\begin{figure*} \\vspace{14cm} \\special{psfile=figmap.ps vscale=95 hscale=95 voffset=-175 hoffset=-25} \\caption{Map of the observed field, taken from our $V,B-V$ photometry. North is up and East left.} \\label{fig-map} \\end{figure*} ", "conclusions": "Although the fits of observed and simulated diagrams are not as satisfying for Be21 as they have been for other systems examined with the same method, we have been able to determine a fairly consistent confidence interval for its distance, age, reddening and metallicity (see Table 5) by selecting the most reliable among the models described in the previous section. They place it among the old metal poor open clusters, in a region far from the Galactic centre and of moderately high reddening. \\subsection{Distance and reddening} We have derived a distance slightly smaller than previous studies, while our evaluation of the reddening is fairly consistent with past works. No previous indication of differential reddening was given in the literature, but our data definitely show it. JP94 found, for the red clumps of 23 open clusters with \\dv$\\ge$1.0 (i.e. older than about 1.5 Gyr) a mean absolute magnitude $M_V=0.9 \\pm 0.4$, and a mean intrinsic colour $(B-V) = 0.95 \\pm 0.10$. In our case, these mean values, when applied to the observed $V$=16.80 and $B-V$=1.55, would imply $(m-M)_V$=15.90, and \\ebv=0.60, or $(m-M)_0$=14.04. From our best simulations, we obtain instead $(m-M)_0 \\simeq $ 13.5, \\ebv=0.76, corresponding to $(m-M)_V$=15.86. In other words, the clump-based distances seem to agree, but the colour of the clump stars seems to be quite different from the mean. Part of this discrepancy may be due to the high reddening affecting Be21, whose differential effect on blue and red stars leads to an apparent shrink by 0.04 of the true colour difference between TO and clump stars (Fernie 1963, Twarog 1998 private communication). On the other side, TAAT97, derive a mean $M_V=0.6 \\pm 0.1$ for ten clusters not too metal-rich ranging from NGC7789 to Mel66, i.e. approximately from 1 to 5 Gyr, for which they try to measure the distance in a quite homogenous way. This translates, in our case, to $(m-M)_V$=16.2; since they do not cite a mean intrinsic $(B-V)$ no further comparison with our best choice for the reddening is possible. There are no completely reliable reddening determination for this cluster since the UBV data of CJ do not reach the MS, but our determination and that by Janes (1991) agree well. We have further compared our finding with what is expected from the spatial distribution of interstellar extinction near the Galactic plane. To this end we have considered the studies of FitzGerald (1968, fig. 3h) and Neckel \\& Klare (1980, fig. 6i). In both cases, a reasonable estimate deduced from their data for low Galactic latitudes and the right longitude, is \\ebv $\\sim$ 0.8; FitzGerald's (1968) observations also allow for a lower value, closer to 0.5, but seem to exclude the high values, close to 1, needed by the lower metallicity tracks of any group. Janes (1991) and JPM94 give a distance modulus \\mmm=13.9$\\pm$0.2, somewhat larger than our results for every set of tracks. Carraro et al. (1998), working on the same data, cite a Galactocentric distance of 14.5 Kpc, also implying a distance modulus \\mmm$\\simeq$13.9. We have no good explanation for this difference, but we must emphasize that adopting \\mmm=13.9 we would be forced to select younger ages, and this would have two major drawbacks: a worse disagreement with literature ages (see next), and a worse reproduction of the MS shape in the synthetic CMDs. \\subsection{Age} Also in the case of the age, we seem to have found a value lower than given in literature. We can explain the discrepancy partly by the different techniques adopted, partly by the better quality of our data. The various parameter combinations all converge to a fairly small range of possible ages (2.1 to 2.8 Gyr, with favorite age around 2.2 Gyr). In fact, the only largely discrepant value found in our analysis is for the FRANEC98 Z=0.01 tracks, a value in strong disagreement with the spectroscopically determined metallicity. Despite the uncertainties involved in the age determination with our method, we consider it still more reliable than ages derived by other means. Nonetheless, it is not always feasible to determine the age of a cluster with the proper method of synthetic diagram fitting: to do so, high quality data, both deep and precise, are needed, and the process itself is complex. To apply this technique to all the objects of interest takes a long time, while the properties of the whole sample of open clusters are needed to study the Disc population and evolution. For this reason, several parameterizations of cluster ages, based on a handful of well studied objects, have been proposed. These methods, if uncertain in absolute value for the single cluster, yield a reasonable age ranking for the cluster system. These parameterizations are usually based on a difference, in magnitude and/or in colour, between well recognizable points of the CMD (usually the TO and the red clump), as this is much easier to measure than any absolute quantity. Note though that the precise definition of the two points, and especially of the TO, changes among authors. We will cite here the three following examples: i) Anthony-Twarog \\& Twarog (1985 and later works) use the magnitude difference between the red giant clump and {\\it the brightest point at the TO} ($\\delta V_T$) coupled with the difference in colour between the red giant branch at the position of the clump and the bluest point of the TO ($\\delta(B-V)_T$); ii) JP94 use a similar \\dv, but measured between the red clump and {\\it the inflection point between the MS TO and the base of the giant branch}; iii) CC94 define their $\\Delta V$ as JP94, but assume that {\\it the reference TO luminosity is 0.25 mag fainter than observed in the CMD}, to take into account the fact that presence of unresolved binaries tends to brighten the TO region. In the case of Be21, we have: \\dv=1.8 (our measure), $\\delta V_T$=1.2, \\dv(JP94)=1.6, $\\Delta V$=1.55. All these different definitions try to circumvent the problem that the TO point is not always easily identified in open clusters, due to field stars and binaries contamination and/or paucity of stars. The strength of our kind of analysis is that we do not judge on the basis of the observed CMD alone: we know from the tracks the exact location of the TO in each of our simulated CMDs. Given this, we have chosen to measure the magnitude difference as defined above at what we believe to be the true TO, i.e. at the point corresponding to the hottest MS point in the evolutionary tracks. JP94 correlated \\dv ~with cluster ages. The calibration of their Morphological Age Index (MAI, expressed in Gyr) translates for Be21 to an age of 2.8 Gyr (based upon \\dv=1.6, PJM94), marginally inconsistent with what we get from the direct comparison with evolutionary tracks (see Section 4). The age difference does not arise from any discrepancy in the two sets of data: we find \\dv=1.8, quite consistent with the value given by PJM94 considering that we measure it in a slightly different way. However, we have found in the past (e.g. in the case of NGC2506, Marconi et al. 1997) that the MAI tends to overestimate ages. Anthony-Twarog \\& Twarog (1985, revised by Twarog \\& Anthony-Twarog 1989) proposed the so called Morphological Age Ratio (MAR), defined as MAR = $\\delta V_T / \\delta (B-V)_T$ (see above). This index is independent of reddening and almost independent of metallicity (Anthony-Twarog \\& Twarog 1985, Buonanno et al. 1989). The calibration of the relation between MAR and ages has changed from: age = 1.4$\\times$MAR Gyr (Anthony-Twarog \\& Twarog 1985) to: age = 2.0$\\times$MAR Gyr (Twarog \\& Anthony-Twarog 1989). Applying their definition to our CMD, we find the values given in Table 6, and an age of about 2.7 to 3.9 Gyr, depending on the adopted calibration. With no attempt to give a new calibration of the MAR-age relation, simply adopting for the clusters we studied the parameters and ages in Table 6, we find: age = 2.3 $\\times$ MAR -- 2.6 ($r.m.s.$ = 0.9). We did not include Be17 in this computation: it represents an extreme of the interpolation and we felt that the parameters derived from the published diagrams were too insecure. This relation gives for Be21 an age of 1.9 Gyr. Carraro et al. (1998) derive for Be21 an age of 3.1 Gyr, based on the PJM94 data and the synthetic CMD method using the Padova tracks. The difference with our results, obtained employing the same sets of tracks (although we do not interpolate in metallicity as they do), may perhaps be explained simply with the worse quality of the observational data they use. Certainly, in no case are we able to reach self-consistently such a large age. \\subsection{Metallicity} We note that our method is unable to solve the problem of the cluster precise metal abundance. The comparison with evolutionary tracks can only give a coarse indication of metallicity. Too many variables are present in tracks computation to discriminate metallicity to such an extent. In fact, tracks nominally closer to the metallicity derived for Be21 from spectroscopic measurements ([Fe/H]=--0.97, or Z $\\simeq$ 0.002, FJ93) appear less consistent with the observed CMD than tracks more metal rich, because of the large colour extent of the subgiant branch and, in some cases, of an excessively high reddening required to reproduce the observed colour of the MS. Anyway, what can be said is that the best fits are obtained for the slightly more metal-rich combinations, i.e. for Z=0.006 (\\fe $\\simeq$ -0.5) or 0.004 (\\fe $\\simeq$ -0.7) as compared to 0.001 (\\fe $\\simeq$ -1.3). This would go in favour of a metal abundance slightly higher than measured by FJ93. There is the possibility that the FJ93 scale may be underestimating cluster metallicities. TAAT97 compared it with abundances based on DDO photometry and found the FJ93 values systematically low. Another example may be the couple of clusters examined by Gratton \\& Contarini (1994): they observed two giants in each cluster at high-resolution and high S/N (R=30,000, S/N $\\simeq$ 100) and found for NGC2243 and Mel66 the values [Fe/H]=--0.48 and --0.38 respectively, to be compared with [Fe/H]=--0.56 and --0.51 (FJ93). FJ93 emphasized the fact that the actual value derived for the cluster metallicity from their spectra is strongly dependent on the adopted reddening: the $\\pm$0.2 mag error on reddening in Janes (1991) allows for a formal uncertainty of $\\pm$0.3 dex in metallicity. They also find marginal support for a \\ebv ~value on the higher side, hence for a metal abundance slightly higher than the [Fe/H]=--0.97 they give. This goes in the same direction suggested by our comparisons, even if we do not find any convincing evidence for a larger reddening. Finally, we have identified the four stars studied by FJ93 (N$_{CJ}$ = 50, 406, 413, 415a, which correspond to N$_{our}$ = 50, 67, 20, 51 respectively), to check if they may be influenced by the differential reddening we found; but we consider it quite improbable, since all the four objects are within 1 arcmin from the centre of the cluster. No conclusive word can be said on Be21 metallicity, which would instead be important to know with high precision, since it could represent the lowest value for the open clusters in our Galaxy. A decisive answer would come from high resolution spectroscopy coupled with fine abundance analysis on the four stars examined by FJ93, already known to be cluster members. \\bigskip\\bigskip\\noindent ACKNOWLEDGEMENTS \\noindent We warmly thank A. Chieffi, M. Limongi and, specially, O. Straniero for having not only distributed the new FRANEC tracks in advance of publication, but even in format suitable for our purposes. We also thank J.C. Mermilliod for kindly making available his invaluable BDA open clusters database and for useful comments. We are grateful to the referee (Bruce Twarog) for his comments, extremely useful both to improve the clarity of the paper and for future applications. The bulk of the numerical code for CMD simulations has been provided by Laura Greggio. This research has made use of the Simbad database, operated at CDS, Strasbourg, France." }, "9804/astro-ph9804137_arXiv.txt": { "abstract": "We analyze the data of low--energy cosmic--ray $\\bar p$ spectrum, recently published by the BESS Collaboration, in terms of newly calculated fluxes for secondary antiprotons and for a possible contribution of an exotic signal due to neutralino annihilation in the galactic halo. We single out the relevant supersymmetric configurations and discuss their explorability with experiments of direct search for particle dark matter and at accelerators. We discuss how future measurements with the Alpha Magnetic Spectrometer (AMS) on the Shuttle flight may disentangle the possible neutralino--induced contribution from the secondary one. ", "introduction": "A recent analysis \\cite{bess95} of the data collected by the balloon--borne BESS spectrometer on cosmic--ray antiprotons during its flight in 1995 (hereafter referred to as BESS95 data) has provided the most detailed information on the low--energy cosmic--ray $\\bar{p}$'s spectrum currently available: 43 antiprotons have been detected, grouped in 5 narrow energy windows over the total kinetic--energy range $180 ~ {\\rm MeV} \\leq T_{\\bar{p}} \\leq 1.4 ~{\\rm GeV}$. With this experiment the total number of measured cosmic--ray antiprotons in balloon--borne detectors over a period of more than 20 years \\cite{all,hof,mitchell,moiseev,barbiellini} has more than doubled. Most remarkably, the BESS95 data provide a very useful information over the low--energy part of the ${\\bar p}$ flux, where a possible distortion of the spectrum expected for secondary $\\bar{p}$'s (i.e., antiprotons created by interactions of primary cosmic--ray nuclei with the interstellar medium) may reveal the existence of cosmic--ray antiprotons of exotic origin (for instance, due to pair annihilation of relic particles in the galactic halo \\cite{th,noi,mitsui}, to evaporation of primordial black holes \\cite{mitsui,kww} or to cosmic strings \\cite{strings}). In fact, a possible discrimination between primary (exotic) and secondary $\\bar p$'s is based on the different features of their low--energy spectra: in this energy regime ($ T_{\\bar p} \\lsim $ 1 GeV) interstellar (IS) secondary $\\bar p$ spectrum is expected to drop off very markedly because of kinematical reasons \\cite{gaisser}, while exotic antiprotons show a milder fall off. However, as will be discussed later on, this discrimination power is somewhat hindered by solar modulation and by some other effects affecting particle diffusion in the Galaxy. In Fig. 1 we report the cosmic--ray $\\bar p$ flux at the top of the atmosphere (hereafter referred to as TOA flux) measured by BESS95 \\cite{bess95}. For experimental data referring to other measurements with much less statistics see Refs.\\cite{all,hof,mitchell,moiseev,barbiellini}. Also displayed in Fig. 1 are the minimal, median and maximal fluxes expected for secondary antiprotons at the time of the BESS95 data taking. These fluxes have been derived with a procedure which is described in detail in Secs. II--V. A comparison of the BESS95 data with the theoretically expected fluxes for secondary $\\bar p$'s, as displayed in Fig. 1, leads to the following considerations: i) the experimental data are consistent with the theoretically expected secondary flux, within the experimental errors and the theoretical uncertainties; however, ii) the experimental flux seems to be suggestive of a flatter behaviour, as compared to the one expected for secondaries ${\\bar p}$'s. Thus, natural questions arise, such as: a) how much room for exotic ${\\bar p}$'s would there be in the BESS95 data, for instance in case the secondary flux is approximately given by the median estimate of Fig. 1, b) how consistent with the current theoretical models would be the interpretation of the BESS95 data in terms of a fractional presence of exotic antiprotons, and c) how this interpretation might be checked by means of independent experiments? In the present note we address these questions within an interpretation of a possible excess of $\\bar p$'s at low energies in terms of primary antiprotons generated by relic neutralinos in the galactic halo \\cite{note1}. The present analysis \\cite{note2} is mostly meant to a clarification of many theoretical points which will be even more crucial, when a much more statistically significant experimental information on low--energy cosmic--ray antiprotons will be made available by forthcoming experiments: AMS on the precursor Shuttle flight in May 1998 and on the International Space Station Alpha (ISSA) in January 2002 \\cite{ams}, the satellite--borne PAMELA experiment \\cite{pamela} and balloon--borne measurements \\cite{spill}. Our paper is organized as follows. In Sec. II we discuss the cosmic--ray IS proton spectrum which will be subsequently employed in deriving the secondary antiprotons. In the same section we also illustrate how we treat the solar modulation to connect the IS spectra to the corresponding TOA fluxes. In Sec. III we discuss the sources of cosmic antiprotons, both of primary and of secondary origins. Cosmic rays diffusion properties are derived in Sec. IV; the TOA $\\bar p$ spectra are given in Sec. V. In Sec. VI we compare our theoretical fluxes with the BESS95 data and single out the neutralino configurations which may be relevant for the present problem. Secs. VII and VIII are devoted to an analysis on how these supersymmetric configurations can be explored by direct searches for relic neutralinos and by experimental investigation at accelerators. Conclusions and perspectives in terms of the forthcoming measurements of low--energy cosmic--ray $\\bar p$'s are illustrated in Sec. IX. ", "conclusions": "We have presented a new analysis of the cosmic--ray antiprotons flux, expected on the basis of secondary $\\bar p$'s, generated by interactions of cosmic--ray primaries with the interstellar medium, and of a possible exotic primary source of $\\bar p$'s, originated by neutralino--neutralino annihilations in the Galactic halo. Improvements over previous calculations of secondaries depend mostly on: i) the use of a two--zone propagation model for diffusion of cosmic rays in the halo instead of the standard leaky box model; ii) the inclusion of an energy--loss effect in the propagation properties of cosmic rays (important for the antiproton low energy range considered in this paper); iii) the use of the new data on primary cosmic--ray proton spectrum, as measured by IMAX \\cite{imaxp} and CAPRICE \\cite{caprice}. The neutralino--induced $\\bar p$ flux has been evaluated in a MSSM at the electroweak scale, which incorporates all current accelerator constraints. Use of supergravity--inspired unification conditions at large energy scale has been avoided in order not to arbitrarily constrain the neutralino phenomenology \\cite{bere}. Solar modulation of the antiproton flux has been improved by analyzing the most complete set of data over the solar cycles \\cite{papini} and the data on the proton spectrum of Refs. \\cite{imaxp,caprice}. We have found that the most statistically relevant data on cosmic--ray antiprotons at low--energy \\cite{bess95} leave some room for a possible signal from neutralino annihilation in the galactic halo. We have discussed how the relevant supersymmetric configurations may be explored with direct experiments for particle dark matter search and at accelerators. We have shown how the interplay between measurements of cosmic--ray $\\bar p$'s and direct search experiments for relic particles is very intriguing and quite important in view of the significant improvements expected in these two classes of experiments in the near future. The present analysis stresses the great interest for the forthcoming AMS measurements with the Shuttle flight and on the ISSA \\cite{ams}, as well as for other future measurements with balloon--borne experiments (IMAX\\cite{mitchell}, BESS\\cite{moiseev}) and with satellites (PAMELA) \\cite{pamela}, for disentangling the secondary $\\bar p$ flux from a possible primary signal of exotic nature. As an example, we give in Fig. 18 the distribution of measurements expected for AMS with the Shuttle flight according to two different hypothesis: a) dominance of the secondary contribution (lower sequence of crosses), b) significant contribution due a neutralino--induced signal (upper sequence of crosses). In our evaluation of the expected measurements we have taken into account geomagnetic cut--off effects and the expected AMS overall acceptance \\cite{batt}. \\begin{center} {\\bf Acknowledgments} \\end{center} We thank Roberto Battiston and Aldo Morselli for interesting discussions about experimental aspects related to the present paper. P.S. would like to thank the French Programme National de Cosmologie for its support. \\newpage \\begin{table} \\caption{Values of the parameters in the expressions (\\ref{eq:energy}) and (\\ref{eq:rigidity}) for the IS proton flux and of the solar--modulation parameter $\\Delta$. These values are obtained by best--fitting the data of Refs.[19-20] with Eqs. (\\ref{eq:energy}) and (\\ref{eq:rigidity}), either over the entire energy range or only over the high--energy ($T_p \\geq 20$ GeV) range. First and third sets of values refer to 3--parameters fits (with Eqs.(\\ref{eq:energy}) and (\\ref{eq:rigidity}), respectively), second and fourth sets refer to 2--parameters fits at fixed $\\Delta$, (with Eqs.(\\ref{eq:energy}) and (\\ref{eq:rigidity}), respectively). } \\begin{center} \\begin{tabular}{|c|c|c|c|} \\hline & IMAX & CAPRICE & Comments \\\\ \\hline \\hline A & 12,300$\\pm$1,700 & 17,600$\\pm$500 & entire energy\\\\ $\\alpha$ & 2.67$\\pm$0.03 & 2.81$\\pm$0.01 & range \\\\ $\\Delta$ & 510$\\pm$40 & 390$\\pm$ 5 & \\\\ \\hline A & 12,300$\\pm$3,000 & 19,600$\\pm$ 3,000 & \\\\ $\\alpha$ & 2.67$\\pm$0.06 & 2.85$\\pm$0.04 & $T_p \\geq$ 20 GeV \\\\ $\\Delta$ & 510 (fixed) & 390 (fixed) & \\\\ \\hline B & 16,200$\\pm$2,000 & 26,000$\\pm$1,200 & entire energy\\\\ $\\gamma$ & 2.73$\\pm$0.03 & 2.91$\\pm$0.02 & range \\\\ $\\Delta$ & 795$\\pm$35 & 710$\\pm$10 & \\\\ \\hline B & 13,700$\\pm$4,100 & 22,800$\\pm$ 3,700 & \\\\ $\\gamma$ & 2.69$\\pm$0.06 & 2.88$\\pm$0.04 & $T_p \\geq$ 20 GeV \\\\ $\\Delta$ & 795 (fixed) & 710 (fixed) & \\\\ \\hline \\end{tabular} \\end{center} \\end{table} \\vfill \\eject \\newpage \\begin{center} {\\bf Figure Captions} \\end{center} {\\bf Fig. 1} - TOA antiproton flux as a function of the antiproton kinetic energy. The experimental points are the BESS95 data \\cite{bess95}. The curves are the median (solid line), minimal (dotted line) and maximal (dashed line) secondary fluxes calculated in this paper, solar--modulated at the time of the BESS95 measurement. {\\bf Fig. 2} - TOA spectra of IMAX (full circles) \\cite{imaxp} and of CAPRICE (open circles) \\cite{caprice} with our best--fit curves with parametrization of Eq. (\\ref{eq:energy}). (The error bars are not shown when they are smaller than the dimension of the circles.) {\\bf Fig. 3} - TOA spectra of IMAX (full circles) \\cite{imaxp} and of CAPRICE (open circles) \\cite{caprice}. The solid, dotted and dashed lines denote the median, minimal and maximal IS proton fluxes, respectively, as discussed in Sec. II. (The error bars are not shown when they are smaller than the dimension of the circles.) {\\bf Fig. 4} - The grammage $\\Lambda_e$ of the CNO primary elements (solid) as inferred from a two--zone diffusion model of the propagation of cosmic rays in the Galaxy. It is plotted as a function of the kinetic energy per nucleon. The dashed curve features the grammage corresponding to protons while the dotted lines delineate the interval of escape lengths inferred from the Ficenec {\\it et al.} \\cite{Ficenec} observations on $^{3}$He at TOA energies comprised between 100 MeV/n and 1.6 GeV/n. \\label{fig:grammage} {\\bf Fig. 5} - IS secondary antiproton spectra as functions of the ${\\bar p}$ kinetic energy. Solid, dotted and dashed lines denote the fluxes obtained from the median, minimal and maximal IS primary proton fluxes. The dot--dashed line denotes the median ${\\bar p}$ spectrum, when the ${\\bar p}$ energy losses are neglected. {\\bf Fig. 6} - Coefficient $C_{\\rm susy}(T_{\\bar p},f)$ as a function of the ${\\bar p}$ kinetic energy for different values of the flattening parameter $f$. {\\bf Fig. 7} - Time variation of the solar--modulation parameter $\\Delta$. Full circles represent the best--fit values to the PGS average fluxes at minima (MIN) and at maxima (MAX) and to the fluxes of IMAX \\cite{imaxp} and of CAPRICE \\cite{caprice}; the open circle refers to the BESS95 data taking period and the cross denotes the extrapolated value of $\\Delta$ at the time relevant for the future AMS Shuttle flight (May 1998). {\\bf Fig. 8} - Solar modulation of the IS median secondary antiproton flux calculated in this paper. Solid line is the IS spectrum; dashed (dotted) line is the solar--modulated spectrum at minima (maxima). {\\bf Fig. 9} - Solar modulation of the IS antiproton flux, due to neutralino annihilation for the representative neutralino configuration with $m_{\\chi} = 62$ GeV, $P = 0.98$ and $\\Omega_{\\chi} h^2 = 0.11$. Solid line is the IS spectrum; dashed (dotted) line is the solar--modulated spectrum at minima (maxima). {\\bf Fig. 10} - TOA antiproton fluxes versus the antiproton kinetic energy. The BESS95 data \\cite{bess95} are shown by crosses. The dashed line denotes the median secondary flux, the dotted one denotes the primary flux due to neutralino annihilation in the halo for a neutralino configuration with $m_{\\chi} = 62$ GeV, $P = 0.98$ and $\\Omega_{\\chi} h^2 = 0.11$. Solid line denotes the calculated total flux. {\\bf Fig. 11} - Scatter plots for configurations of set $M$ (a) and set $N$ (b) in the P--$\\tan \\beta$ plane. {\\bf Fig. 12} - Scatter plots for configurations of set $M$ (a) and set $N$ (b) in the P--$m_{\\chi}$ plane. {\\bf Fig. 13} - Scatter plots for configurations of set $R$ in the P--$\\tan \\beta$ plane (a) and in the P--$m_{\\chi}$ plane (b). {\\bf Fig. 14} - Scatter plots for configurations of set $R$ in the P--$\\tan \\beta$ plane (a) and in the P--$m_{\\chi}$ plane (b) for a flattening of $f = 0.5$. {\\bf Fig. 15} - Scatter plot of the values of $\\xi \\sigma_{\\rm scalar}^{\\rm nucleon}$ versus the neutralino mass for the configurations of set $M$ (a) and of set $N$ (b). The open curve denotes the (90 \\% C.L.) upper bound obtained from experimental data of Ref. \\cite{few}. The region delimited by a closed contour is the one singled out by the experiment of Ref. \\cite{dama} as possibly indicative of an annual modulation effect. The total local dark matter density is normalized here to the value $\\rho_l = 0.4$ GeV cm$^{-3}$. The dashed line shows the discovery potential in case of an improvement by a factor of 10 in current sensitivities for experiments of direct search for particle dark matter. {\\bf Fig. 16} - Correlation between $\\xi \\sigma_{\\rm scalar}^{\\rm nucleon}$ and the neutralino relic density $\\Omega_{\\chi} h^2$ for configurations of set $M$. {\\bf Fig. 17} - Scatter plot for configurations of set $M$ in the $m_h$--$\\tan \\beta$ plane. The region on the left of the dashed line denoted by (a) is excluded by current LEP experimental data \\cite{lep182}, the one on the right of the dashed line (b) is theoretically disallowed. The other lines display the LEP reach at luminosity $L = 200$ pb$^{-1}$ and various energies \\cite{alt}: (A) discovery potential at ${\\sqrt s} = 192$ GeV; (B) discovery potential at ${\\sqrt s} = 200$ GeV; (C) exclusion at ${\\sqrt s} = 200$ GeV. {\\bf Fig. 18} - Expected distribution of measurements with the AMS Shuttle flight according to two different hypothesis: a) dominance of the secondary contribution (lower sequence of crosses), b) significant contribution due a neutralino--induced signal (upper sequence of crosses). The dashed line denotes the secondary flux, the dotted one denotes the primary flux due to neutralino annihilation in the halo for a neutralino configuration with the representative values: $m_{\\chi} = 62$ GeV, $P = 0.98$ and $\\Omega_{\\chi} h^2 = 0.11$. Solid line denotes the calculated total flux." }, "9804/astro-ph9804301_arXiv.txt": { "abstract": "We present a fully covariant and gauge-invariant calculation of the evolution of anisotropies in the Cosmic Microwave Background (CMB) radiation. We use the physically appealing covariant approach to cosmological perturbations, which ensures that all variables are gauge-invariant and have a clear physical interpretation. We derive the complete set of frame-independent linearised equations describing the (Boltzmann) evolution of anisotropy and inhomogeneity in an almost Friedmann-Robertson-Walker (FRW) Cold Dark Matter (CDM) universe. These equations include the contributions of scalar, vector and tensor modes in a unified manner. Frame-independent equations for scalar and tensor perturbations, which are valid for any value of the background curvature, are obtained straightforwardly from the complete set of equations. We discuss the scalar equations in detail, including the integral solution and relation with the line of sight approach, analytic solutions in the early radiation dominated era, and the numerical solution in the standard CDM model. Our results confirm those obtained by other groups, who have worked carefully with non-covariant methods in specific gauges, but are derived here in a completely transparent fashion. ", "introduction": "The cosmic microwave background radiation (CMB) occupies a central role in modern cosmology. It provides us with a unique record of conditions along our past lightcone back to the epoch of decoupling (last scattering), when the optical depth to Thomson scattering rises suddenly due to Hydrogen recombination. Accurate observations of the CMB anisotropy should allow us to distinguish between models of structure formation and, in the case of non-seeded models, to infer the spectrum of initial perturbations in the early universe. Essential to this programme is the accurate and reliable calculation of the anisotropy predicted in viable cosmological models. Such calculations have a long history, beginning with Sachs \\&\\ Wolfe (1967) who investigated the anisotropy on large scales $(\\gtrsim 1^{\\circ})$ by calculating the redshift back to last scattering along null geodesics in a perturbed universe. On smaller angular scales one must address the detailed local processes occurring in the electron/baryon plasma prior to recombination, and the effects of non-instantaneous last scattering. These processes, which give rise to a wealth of structure in the CMB power spectrum on intermediate scales and damping on small scales (see, for example, Silk (1967, 1968)), are best addressed by following the photon distribution function directly from an early epoch in the history of the universe to the current point of observation. This requires a numerical integration of the Boltzmann equation, and has been carried out by many groups, of which Peebles \\&\\ Yu (1970), Bond \\&\\ Efstathiou (1984, 1987), Hu \\&\\ Sugiyama (1995), Ma \\&\\ Bertshcinger (1995), Seljak \\&\\ Zaldarriaga (1996) is a representative sample. The calculation of CMB anisotropies is simple in principle, but in reality is plagued with subtle gauge issues (Stoeger, Ellis, \\&\\ Schmidt 1991; Stoeger et al.\\ 1995; Challinor \\&\\ Lasenby 1998). These problems arise because of the gauge-freedom in specifying a map $\\Phi$ between the real universe (denoted by $S$) and the unperturbed background model(denoted by $\\bar{S}$)~\\cite{ellis89a}, which is usually taken to be a Friedmann-Robertson-Walker (FRW) universe. The map $\\Phi$ identifies points in the real universe with points in the background model, thus defining the perturbation in any quantity of interest. For example, for the density $\\rho$ as measured by some physically defined observer, the perturbation at $x\\in S$ is defined to be $\\delta \\rho(x)\\equiv \\rho(x) - \\bar{\\rho}(\\bar{x})$, where $\\bar{\\rho}$ is the equivalent density in the background model, and $x$ maps to $\\bar{x}$ under $\\Phi$. The map $\\Phi$ is usually (partially) specified by imposing coordinate conditions in $S$ and $\\bar{S}$. Any residual freedom in the map $\\Phi$ after the imposition of the coordinate conditions (gauge-fixing) gives rise to the following gauge problems: (i) the map cannot be reconstructed from observations in $S$ alone, so that quantities such as the density perturbation, which depend on the specific map $\\Phi$, are necessarily not observable; (ii) if the residual gauge freedom allows points in $\\bar{S}$ to be mapped to physically inequivalent points in $\\bar{S}$ in the limit that $S=\\bar{S}$, then unphysical gauge mode solutions to the linearised perturbation equations will exist. There are several ways to deal with the gauge problems described above. In the earliest approach~\\cite{lifshitz46}, one retains the residual gauge freedom (in the synchronous gauge) but keeps track of it so that gauge mode solutions can be eliminated. Furthermore, the final results of such a calculation must be expressed in terms of the physically relevant, gauge-invariant quantities. Although there is nothing fundamentally wrong with this approach if carried out correctly, it suffers from a long history littered with confusion and errors. The need to express results in terms of gauge-invariant variables suggests that it might be beneficial to employ such variables all along as the dynamical degrees of freedom in the calculation. A further advantage of such an approach is that gauge modes are automatically eliminated from the perturbation equations when expressed in terms of gauge-invariant variables. This is the approach adopted by Bardeen (1980), who showed how to construct gauge-invariant variables for scalar, vector and tensor modes in linearised perturbation theory, by taking suitable linear combinations of the gauge-dependent perturbations in the metric and matter variables. This approach has been used in several calculations of the CMB anisotropy (see, for example, Abbott \\&\\ Schaefer (1986) and Panek (1986)). However, the Bardeen variables are not entirely satisfactory. The approach is inherently linear, so that the variables are only defined for small departures from FRW symmetry. Furthermore, the approach assumes a non-local decomposition of the perturbations into scalar, vector and tensor modes at the outset, each of which is then treated independently. As a result, the Bardeen variables are only gauge-invariant for the restricted class of gauge-transformations that respect the scalar, vector and tensor splitting. Finally, although the Bardeen variables are gauge-invariant, they are not physically transparent, in that, in a general gauge, they do not characterise the perturbations in a manner that is amenable to simple physical interpretation. An alternative scheme for the gauge-invariant treatment of cosmological perturbations was given by Ellis \\&\\ Bruni (1989) (see also, Ellis, Hwang, \\&\\ Bruni (1989)) who built upon earlier work by Hawking (1966). In this approach, which is derived from the covariant approach to cosmology/hydrodynamics of Ehlers and Ellis (Ehlers 1993; Ellis 1971), the perturbations are described by gauge-invariant variables that are covariantly defined in the real universe. This ensures that the variables have simple physical interpretations in terms of the inhomogeneity and anisotropy of the universe. Since the definition of the covariant variables does not assume any linearisation, exact equations can be found for their evolution, which can then be linearised around the chosen background model. Furthermore, the covariant approach does not employ the non-local decomposition into scalar, vector or tensor modes, at a fundamental level. If required, the decomposition can be performed at a late stage in the calculation to aid solving the equations. Even if one denies that working with gauge-invariant variables is a significant advantage, the key advantage of the covariant approach, however, is that one is able to work exclusively with physically relevant quantities, satisfying equations that make manifest their physical consequences. The covariant and gauge-invariant approach has already been applied to the line of sight calculation of CMB anisotropies under the instantaneous recombination approximation (Dunsby 1997; Challinor \\&\\ Lasenby 1998), and has been used to obtain model-independent limits on the inhomogeneity and anisotropy from measurements of the CMB anisotropy on large scales~\\cite{maartens95}. In this paper, we extend the methodology developed in these earlier papers, to give a full kinetic theory calculation of CMB anisotropies valid on all angular scales. Our motivation for reconsidering this problem is two-fold. Firstly, it is our belief that the covariant and gauge-invariant description of cosmological perturbations provides a powerful set of tools for the formulation of the basic perturbation equations, and their subsequent interpretation, which are superior to the techniques usually employed in such calculations for the reasons discussed above. In particular, by applying covariant methods for the problem of the generation of CMB anisotropies, we can expect the same advantages of physical clarity and unification that have already been demonstrated in other areas, (Ellis et al.\\ 1989; Bruni, Ellis, \\&\\ Dunsby 1992; Dunsby, Bruni, \\&\\ Ellis 1992; Dunsby, Bassett, \\&\\ Ellis 1996; Tsagas \\&\\ Barrow 1997). The approach described here brings the underlying physics to the fore, and can only help to consolidate our rapidly growing understanding of the physics of CMB anisotropies. Furthermore, although we only consider the linearised calculation here, the extension of these methods to the full non-linear case is quite straightforward (Maartens, Gebbie, \\&\\ Ellis 1998). Our second motivation is to perform an independent verification of the results of other groups (for example, Ma \\&\\ Bertschinger (1995)), with a methodology that is free from any of the gauge ambiguities that have caused problems and confusion in the past. Given the potential impact on cosmology of the next generation of CMB data, we believe that the above comments provide ample justification for reconsidering this problem. For definiteness we consider the cold dark matter (CDM) model, although the methods we describe are straightforward to extend to other models. We have endeavoured to make this paper reasonably self-contained, so we begin with a brief overview of the covariant approach to cosmology and define the key variables we use to characterise the perturbations in Section~\\ref{sec_cov}. We then go on to present a complete set of frame-independent equations describing the evolution of the matter components and radiation in Section~\\ref{sec_eqs} in an almost-FRW universe (with arbitrary spatial curvature). These equations, which employ only covariantly defined, gauge-invariant variables, are independent of any harmonic analysis; they describe scalar, vector and tensor perturbations in a unified manner. Many of the equations have simple Newtonian analogues, and their physical consequences are far more transparent than the equations that underlie the metric-based approaches. Equations pertinent to scalar modes, see Section~\\ref{sec_scal}, and tensor modes, see Section~\\ref{sec_tens}, can be obtained from the full set of equations with very little effort, and are useful at this late stage in the calculation as an aid to solving the linearised equations. A significant feature of this approach is that a covariant angular decomposition of the distribution functions is made early on in the calculation, before any splitting into scalar, vector and tensor modes. This allows scalar, vector and tensor modes to be treated in a more unified manner. In particular, the azimuthal dependence of the moments of the distribution functions does not have to be put in by hand (after inspection of the azimuthal dependence of the other terms in the Boltzmann equation), as happens in most metric-based calculations. This is particularly significant for tensor modes where the required azimuthal dependence is non-trivial and is different for the two polarisations of gravitational waves. We consider the equations for scalar modes in considerable detail. We present the integral solution of the Boltzmann multipole equations in a $K=0$ almost-FRW universe, and discuss the relation between line of sight methods (which employ lightlike integrations along the lightcone) and the Boltzmann multipole approach (where a timelike integration is performed). We derive analytic solutions for scalar modes in the early radiation dominated universe, that are used as initial conditions for the numerical solution of the scalar equations, the results of which we describe in Section~\\ref{sec_num}. In Section~\\ref{sec_tens} we give a brief discussion of the tensor equations in the covariant approach. The covariant angular decomposition naturally gives rise to a set of variables that describe the temperature anisotropy in a more direct manner than in the conventional metric-based approaches. This is particularly apparent for tensor perturbations, where the CMB power spectrum at a given multipole $l$ is determined by the $l-2$, $l$ and $(l+2)$-th moments of the conventional decomposition of the photon distribution function, which obscures the physical interpretation of these moments. Finally, we end with our conclusions in Section~\\ref{sec_conc}. Ultimately, our results confirm those of other groups (for example, Ma \\&\\ Bertschinger (1995)) who have performed similar calculations by working carefully in specific gauges, but are obtained here with a unified methodology that is more physically transparent and less prone to lead to confusion over subtle gauge effects. We employ standard general relativity and use a $(+---)$ metric signature. Our conventions for the Riemann and Ricci tensors are fixed by $[\\nabla_{a},\\nabla_{b}] u^{c} = -{\\clr_{abd}}^{c} u^{d}$, and $\\clr_{ab} \\equiv {\\clr_{acb}}^{c}$. Round brackets around indices denote symmetrisation on the indices enclosed, and square brackets denote antisymmetrisation. We use units with $c=G=1$ throughout, and a unit of distance of $\\mpc$ for numerical work. ", "conclusions": "\\label{sec_conc} We have shown how the full kinetic-theory calculation of the evolution of CMB anisotropies and density inhomogeneities can be performed in the covariant approach to cosmology (Ehlers 1993; Ellis 1971), using the gauge-invariant perturbation theory of Ellis \\&\\ Bruni (1989). Adopting covariantly-defined, gauge-invariant variables throughout ensured that our discussion avoided the gauge ambiguities that appear in certain gauges, and that all variables had a clear, physical interpretation. We presented a unified set of equations describing the evolution of photon and neutrino anisotropies and cosmological perturbations in the CDM model, which were independent of a decomposition into scalar, vector or tensor modes and the associated harmonic analysis. Although we only considered the case of linear perturbations around an FRW universe here, it is straightforward to extend the approach to include non-linear effects (Maartens et al. 1998), which should allow a physically transparent discussion of second-order effects on the CMB. Indeed, the ease with which one can write down the exact equations for the physically relevant variables is one of the major strengths of the covariant approach. The linear equations describing scalar modes and tensor modes were obtained from the full set of equations in a straightforward and unified manner, highlighting the advantage of having the full equations (independent of the decomposition into scalar, vector and tensor modes) available. For the scalar case, the Boltzmann multipole equations for the moments of the distribution functions obtained here were equivalent to those usually seen in the literature. However, for tensor modes, the covariant approach led naturally to a set of moment variables that more conveniently describe the temperature anisotropy than those usually employed. For scalar modes, we discussed the solution of the perturbation equations in detail, including the integral solution of the Boltzmann multipole equations and the relation between the timelike integrations performed in the multipole approach to calculating CMB anisotropies, and the lightlike integrations of the line of sight approach. The numerical solution of the scalar equations in a $K=0$, almost-FRW, CDM universe were also discussed. Our numerical results provide independent confirmation of those of other groups, (see, for example, Ma \\&\\ Bertschinger (1995) and Seljak \\&\\ Zaldarriaga (1996)), who have obtained their results by employing non-covariant methods in specific gauges. Typically, these methods require one to keep careful track of all residual gauge-freedom, both to enable identification of any gauge-mode solutions, and to ensure that the final results quoted are gauge-invariant (and hence observable). Fortunately, the isotropy of the photon distribution function in an exact FRW universe ensures that the CMB power spectrum, as calculated from the gauge-dependent perturbation to the distribution function, is gauge-invariant for $l\\geq 1$. We hope to have shown the ease with which the covariant approach to cosmology can be applied to the problem of calculating CMB anisotropies. The covariant and gauge-invariant method discussed here frees one from the gauge problems that have caused confusion in the past, and focuses attention on the physically relevant variables in the problem and the underlying physics. Future work in this area will include the discussion of non-linear effects (Maartens at al. 1998), the inclusion of polarisation, and the effects of hot dark matter, all of which can be expected to bring the same advantages of physical clarity and transparency that we hope to have demonstrated here." }, "9804/astro-ph9804184_arXiv.txt": { "abstract": "This paper describes the first optical spectroscopic survey of class I sources (also known as embedded sources and protostars) in the Taurus-Auriga dark cloud. We detect 10 of the 24 known class I sources in the cloud at 5500--9000 \\AA. All detected class I sources have strong H$\\alpha$ emission; most also have strong [O~I] and [S~II] emission. These data -- together with high quality optical spectra of T Tauri stars in the Taurus-Auriga cloud -- demonstrate that forbidden emission lines are stronger and more common in class I sources than in T Tauri stars. Our results also provide a clear discriminant in the frequency of forbidden line emission between weak-emission and classical T Tauri stars. In addition to strong emission lines, three class I sources have prominent TiO absorption bands. The M-type central stars of these sources mingle with optically visible T Tauri stars in the HR diagram and lie somewhat below both the birthline for spherical accretion and the deuterium burning sequence for disc accretion. ", "introduction": "Examining the earliest phases of low mass stellar evolution requires observations of protostars deeply embedded in the dense cores of nearby molecular clouds. These ``class I'' sources (\\cite{lad87}) have blackbody-like spectral energy distributions that peak at wavelengths of 30--100 \\mum~and bolometric luminosities, $L_{\\rm b} \\approx$ 0.1--100 \\lsun~(\\cite{ada87}; \\cite{mye87}; \\cite{wil89}; \\cite{ken90}; \\cite{gre94}). Despite many detailed studies of their circumstellar environments (see, for example, \\cite{ta91a},b; \\cite{and93}; \\cite{ter93}; \\cite{mor92}, 1995; \\cite{bon96}; \\cite{gom97}; \\cite{hog97}; \\cite{whi97}), understanding the stellar physics of these objects has proved elusive. Comparisons of observed bolometric luminosity functions with models is straightforward but controversial (\\cite{wil89}; \\cite{ken90}; \\cite{fl94a}, 1994b). The apparent lack of photospheric features in many objects has led several groups to abandon the HR diagram as a means for testing stellar evolutionary tracks of the youngest stars. These groups have proposed the bolometric temperature (\\cite{mye98} and references therein), the submillimeter flux (\\cite{sar96}), and the visual extinction (\\cite{ada90}) to replace effective temperature and have developed models to place evolving pre--main-sequence stars in their modified evolutionary diagrams. The accuracy of these techniques remains uncertain, because the methods are new and incompletely tested. In this paper, we report an optical spectroscopic survey designed to detect photospheric absorption features from the central stars of class I sources in the Taurus-Auriga cloud. Although the line-of-sight extinction to the central star is large, $A_V \\approx$ 30--60 mag (\\cite{whi97}), large ground-based telescopes can detect optical light scattered off cavities in the infalling envelopes of many objects. Optical data also provide the best measure of spectral types for pre--main-sequence stars. In general, I-band and J-band data are least contaminated by emission from an accretion disc and its associated boundary layer or magnetic accretion column (\\cite{kh90}). However, the very large continuum veiling detected on near-IR spectra of some class I sources (\\cite{cas92}; \\cite{gr96a}, 1996b) favors I-band spectra, because the J-band veiling can be large if the disc extends to the stellar photosphere (\\cite{kh90}; \\cite{ken96}). Finally, optical spectra of class I sources allow an unambiguous comparison with optically brighter T Tauri stars, which have known spectral types in a well-calibrated system (see, for example, Kenyon \\& Hartmann 1995; KH95 hereafter). Our results provide the first optical detection of M-type absorption features in an embedded protostar. We identify TiO absorption bands in three Taurus-Auriga class I sources; one other star may have TiO features and a fifth star may have K-type absorption features. We use optical spectra of T Tauri stars to calibrate the spectral types of class I sources and then construct a complete HR diagram for the Taurus-Auriga cloud. These data, coupled with new evolutionary tracks for protostars accreting from discs and two spectral types derived from near-IR spectra (\\cite{gr96b}), show that class I sources in Taurus-Auriga mingle with T Tauri stars and lie below the birthline in the HR diagram. We also detect strong emission lines on the spectra of all protostars. Forbidden emission from [N~II] and [S~II] is much more common among class I sources than older, optically brighter stars having the same bolometric luminosity. The fluxes of forbidden emission lines also seem stronger among class I sources than other pre--main-sequence stars in the cloud. We find no evidence that the permitted emission lines, such as H$\\alpha$ and He~I, are more common or stronger than in T Tauri stars. These results extend and confirm previous conclusions that jet activity declines as a pre--main-sequence star contracts to the main-sequence. We describe our observations in Sec. 2, explain our results in Sec. 3, and conclude with a brief discussion in Sec. 4. ", "conclusions": "In the previous sections, we have described the first optical spectroscopic survey of class I, embedded sources in a single molecular cloud. We supplemented these data with high quality optical spectra of a representative sample of older and optically brighter T Tauri stars. The combined set of spectra show that the optical spectra of class I sources qualitatively resemble the optical spectra of T Tauri stars. Our analysis further reveals common physical properties and substantial differences between class I sources and T Tauri stars, as summarized below. Our data provide the first indication that the distribution of stellar spectral types among class I sources may not be very different from that of WTTS and CTTS. Of the five class I sources with strong optical continua, one (L1489 IRS) is a continuum + emission source, three are M-type stars, and another (04264+2433) may have an M-type central star. To the best of our knowledge, {\\it these are the first low mass protostars with measured optical spectral types.} This sample is too small for a meaningful comparison with the distribution of spectral types among more evolved pre--main-sequence stars in the cloud. We note, however, that the median spectral type for WTTS and CTTS is K7-M0 and that the frequency of continuum + emission sources is $\\sim$ 5\\%--10\\% (KH95). Published observations indicate other similarities between class I sources and older pre--main-sequence stars in Taurus-Auriga. First, class I sources have the same intrinsic near-IR colors as do CTTS. Whitney \\etal (1997) show that the observed near-IR colors of class I sources can be modeled as a CTTS surrounded by an infalling envelope with an optical extinction, $A_V \\approx$ 30--60 mag. This analysis leads to the conclusion that the radiation from the star and inner disc of a class I source is similar to that of a T Tauri star (see also \\cite{gr96b}; Calvet \\etal 1997 reach a different conclusion). Second, the bolometric luminosity distributions of class I sources, CTTS, and WTTS are indistinguishable (KH95). All three groups of pre--main-sequence stars have median luminosities of $L_{\\rm b} \\approx$ 0.5--0.8 \\lsun. This unexpected result is supported by the positions of class I sources in the HR diagram. Our data show that class I sources have luminosities and effective temperatures very similar to those of CTTS and WTTS in the cloud. These conclusions are surprising, because a class I source should have a larger luminosity once it has accreted nearly all of its final mass, and this luminosity should decline with time as the star approaches the main sequence (see, for example, \\cite{sta83}, 1988; \\cite{pal93}; \\cite{har97}; Fig. 8). The current sample, however, is too small to test stellar models in detail. The errors in luminosity and effective temperature are also too large. Observations with the next generation of ground-based telescopes will undoubtedly expand the sample, reduce the errors, and provide better tests of protostellar accretion theory. One feature that distinguishes class I sources is their strong forbidden-line emission. As a group, class I sources are much more likely to have forbidden-line emission than CTTS or WTTS (Fig. 7). This result confirms previous conclusions from imaging data (e.g., \\cite{gom97}) and indicates that class I sources are more likely to drive outflows than CTTS or WTTS (see also \\cite{bon96}; \\cite{mor92}, 1994). The equivalent widths of the forbidden lines are also larger in class I sources than in CTTS or WTTS. Although some large equivalent widths may be due to very weak optical continua, the [S~II] equivalent widths in HH31 IRS2 -- a class I source with a prominent TiO absorption band -- are larger than observed in {\\it any} CTTS in our sample (see Tables 1--2). Deeper optical spectra of our sample and other class I sources would clarify this point. Our sample is not large enough to test whether class I sources also have more prominent {\\it permitted} emission lines than CTTS. The median H$\\alpha$ equivalent width for class I sources, $\\sim$ 90 \\AA, is much larger than the median equivalent width for CTTS, $\\sim$ 30--40 \\AA. This difference is roughly what we expect if class I sources have larger continuum veiling than CTTS (\\cite{cas96}; \\cite{gre97}) and if the H$\\alpha$ equivalent width correlates with veiling (\\cite{har95} and references therein). However, the frequency of He~I $\\lambda\\lambda$5876, 6678 emission among class I sources is roughly comparable to that among CTTS. We measure a He~I emission frequency of 50\\% among 6 class I sources with reasonable signal-to-noise at 6000 \\AA, 57\\% among 14 flat-spectrum sources, and 65\\% among 46 class II sources. For both emission lines, the class I sample is probably biased against small equivalent widths, because class I sources without emission lines are probably fainter than sources with emission lines. A deeper survey with a larger telescope could enlarge the sample of class I sources with high quality optical spectra. These data would provide a good test for differences in the distribution of H$\\alpha$ equivalent widths between class I sources and CTTS. These results fit into the general picture of Taurus-Auriga class I sources developed in KH95 and in Kenyon \\etal (1990). In this picture, class I sources are envelopes of gas and dust falling into the central star-disc system at rates of a few $\\times~10^{-6}~\\msunyr$ (see also \\cite{ada87}; \\cite{ke93a}, 1993b; \\cite{whi97}). Bell \\& Lin (1994) show that the stable accretion rate through the disc onto the central star is either very low -- $\\lesssim$ a few $\\times~10^{-7}~\\msunyr$ -- or very high -- $\\gtrsim$ a few $\\times~10^{-5}~\\msunyr$ -- compared to the infall rate. The disc spends most of its time in the low accretion rate state; the disc mass then slowly increases with time until it reaches a critical level and evolves to the high accretion rate state. This model explains the low observed luminosities of nearly all class I sources as well as the occasional high luminosity of a source such as L1551 IRS5. Models with time-dependent disc accretion also qualitatively account for the evolution of forbidden and permitted emission lines in pre--main-sequence stars. We expect the time-averaged accretion rate through the disc to decline as the envelope disperses. If the H$\\alpha$ and other permitted emission lines of class I sources form in the accretion region of the inner disc as in CTTS, then the median H$\\alpha$ equivalent width should decline as a pre--main sequence star evolves from a class I source into a CTTS and then into a WTTS. Most models for jet formation link the mass loss rate in the jet to the mass accretion rate through the disc (see, for example, \\cite{cab90}; \\cite{naj94}; \\cite{sh94a}, 1994b), so we expect forbidden emission to decline as well. Explaining the observations of emission line equivalent widths with a quantitative model of a dispersing envelope and evolving disc, however, is currently beyond our reach. Finally, our results further demonstrate the advantages of optical spectra. Recent surveys of larger samples of class I sources using near-IR spectroscopy have yielded only two spectral types each in Taurus-Auriga (\\cite{cas96}; \\cite{gr96b}) and $\\rho$ Oph (\\cite{gr96b}, 1997). Casali \\& Eiroa (1996; see also \\cite{cas92}; \\cite{gr96b}, 1997) conclude that continuum emission from dust in a circumstellar disc or envelope veils photospheric absorption features on near-IR spectra of class I sources. Preliminary results further suggest that this veiling is larger in class I sources than in CTTS or WTTS (\\cite{cas96}; \\cite{gr96b}, 1997). Dust emission is much weaker relative to a normal stellar photosphere at shorter wavelengths, $\\lesssim 1~\\mu$m, so optical spectra may yet provide the best measure of spectral types in class I sources. \\vskip 6ex We thank the staffs of the MMT, Palomar, and Whipple Observatories for assistance with our observations. Fred Chaffee kindly acquired several spectra of the class II sources listed in Table 1. Susan Tokarz reduced the FAST spectra and graciously assisted with the reduction of the MMT and Palomar spectra. We also thank M. Geller, M. G\\'omez, C. Lada, A. Mahdavi, and B. Whitney for advice and comments. The suggestions of an anonymous referee improved our presentation. Observations at the Palomar Observatory were made as part of a continuing collaborative agreement between Palomar Observatory and the Jet Propulsion Laboratory. Portions of this research were supported by the National Aeronautics and Space Administration through grant NAGW-2919 and by the Space Telescope Science Institute through grant GO-06132.01-94A. C.A.T. thanks the Royal Society and the Hungarian Academy of Sciences for an exchange fellowship during the majority of his contribution to this project. \\vfill \\eject" }, "9804/astro-ph9804229_arXiv.txt": { "abstract": "We present and discuss 25 spectra obtained in November 1996, covering all phases of the CAL 87 binary system. These spectra are superior both in signal-to-noise and wavelength coverage to previously published data so that additional spectral features can be measured. Photometry obtained on the same nights is used to confirm the ephemeris and to compare with light curves from previous years. Analysis of the color variation through the orbital cycle has been carried out using archival MACHO data.\\footnote{This paper utilizes public domain data obtained by the MACHO Project, jointly funded by the US Department of Energy through Lawrence Livermore National Laboratory under contract W7405-Eng-48, the National Science Foundation through the Center for Particle Astrophysics of the University of California undercooperative agreement AST-8809616, and the Mount Stromlo and Sidings Springs Observatory by the Bilateral Science Technology and Regional Development.} When a barely resolved red field star is accounted for, there is no ($V-R$)-color variation, even through eclipse. There have been substantial changes in the depth of minimum light since 1988; it has decreased more than 0.5 mag in the last several years. The spectral features and radial velocities are also found to vary not only through the 0.44-day orbit but also over timescales of a year or more. Possible interpretations of these long-term changes are discussed. The 1996 spectra contain phase-modulated Balmer absorption lines not previously seen, apparently arising in gas flowing from the region of the compact star. The changes in emission-line strengths with orbital phase indicate there are azimuthal variations in the accretion disk structures. Radial velocities of several lines give different amplitudes and phasing, making determination of the stellar masses difficult. All solutions for the stellar masses indicate that the companion star is considerably less massive than the degenerate star. The Balmer absorption-line velocities correspond to masses of $\\sim$1.4M$_{\\odot}$ for the degenerate star and $\\sim$0.4M$_{\\odot}$ for the mass donor. However, the strong He II emission lines indicate a much more massive accreting star, with M$_X>$4M$_{\\odot}$. ", "introduction": "CAL 87 has long been known as one of only a small number of luminous X-ray binaries in the Large Magellanic Cloud (Long, Helfand, \\& Grabelsky 1981, Pakull et al.\\ 1988). Its optical spectrum, with He II and H emission lines on a very blue continuum, shows it to be similar to galactic low-mass X-ray binaries. However, its X-ray spectrum reveals CAL 87 to be one of the rare, very luminous (L$_{bol}\\geq10^{38}$ erg s$^{-1}$) supersoft sources (SSS) which have little or no radiation above $\\sim$0.5 keV (e.g. Tr\\\"umper et al.\\ 1991, Greiner 1996). The SSS are widely thought to be binaries in which a white dwarf is undergoing rapid accretion from a more massive companion, leading to steady nuclear burning on the surface of the white dwarf (van den Heuvel et al.\\ 1992). CAL 87 is unique among supersoft sources in having both optical and X-ray eclipses (Callanan et al.\\ 1989, Cowley et al.\\ 1990: CSCH) which provide extra information about the disk structure and in principle help to constrain the stellar masses. In the original spectroscopic data of CSCH, He II 4686\\AA\\ emission was shown to move with K$=$40 km s$^{-1}$ and proper phasing with respect to the eclipse so that it was interpreted as due to orbital motion of the compact star. CSCH concluded that the compact star had a mass $\\ge$6M$_{\\odot}$, and hence these data implied the presence of a black hole. However, optical spectra taken a few years later, combined with velocities from lines in the far UV (Hutchings et al.\\ 1995), showed a quite different behavior. The velocity amplitude was larger and the phasing was very different, indicating that at times the velocities are not entirely due to orbital motion and the line-formation regions change. One problem is that spectroscopic determination of the masses is not entirely straightforward since the nearly edge-on view of the system ($i\\sim78^{\\circ}$) causes complications in the line profiles due to motions in the accretion disk. In this paper we report on a series of spectroscopic data with improved signal-to-noise (S/N) and new, concurrent photometry which we obtained at CTIO in 1996 in order to conduct a more thorough investigation of CAL 87. ", "conclusions": "We have discovered that the photometric eclipse depth has changed over recent years, and this variation may even be cyclic. The increase of light observed at mid-eclipse may be due to a geometrical change in the disk structure, without a significant change in total luminosity -- such as a bright structure above or below the disk plane. Alternatively, the disk may grow in size in its plane, accompanied by extra absorption due to the outflowing material, along lines-of-sight near the center. The emergence of the Balmer absorption in the 1996 spectra suggests the latter scenario. In 1994 November we obtained three optical spectra of CAL 87 which were reported by Hutchings et al.\\ (1995) in their discussion of eight ultraviolet spectra observed by HST in 1995 January. We have remeasured the optical spectra, although they are of considerably lower quality than the new data and only show the strongest lines. In 1994--5 He II 4686\\AA\\ and 1640\\AA\\ emission-line velocities show a very different phasing (maximum velocity near $\\phi\\sim0.9$) and higher amplitude (K$\\sim$70 km s$^{-1}$) compared to both the CSCH data (K=40 km s$^{-1}$) and present data (K$\\sim$30 km s$^{-1}$). The equivalent width of 4686\\AA\\ was lower in the 1994 spectra than in 1996 by a factor of nearly two. In two of the three 1994 optical spectra weak absorption features are visible at H$\\beta$ and H$\\gamma$, similar to those seen in 1996, but the low S/N make the velocities unreliable. Thus, it appears that there are long-term spectral changes that may include significant non-orbital motions. The O VI lines might give a clean measure of orbital motion, as they arise in the inner disk, but they are very weak, resulting in a large velocity scatter. The measured velocity amplitude (K=35 km s$^{-1}$) leads to the same conclusion as CSCH, that the compact star must be massive, with M$_X>$4M$_{\\odot}$. If instead the velocity amplitude of the compact star is shown by the Balmer absorption lines (K=73 km s$^{-1}$), then the resulting mass diagram is almost identical to that of the supersoft binary SMC 13 (Crampton et al.\\ 1997) and lower masses are determined. We have argued above that the measured He II velocity amplitude may be somewhat lower than the actual velocity of the compact star, because the phasing of absorptions will decrease the apparent velocity extremes, while non-orbital motions within the system add at other phases. We point out that the He II velocities in CSCH, when the eclipse was deeper, gave K=40 km s$^{-1}$. The Balmer-line velocities are also likely to contain some non-orbital motions. Thus K=73 km s$^{-1}$ is a reasonable upper limit for the motion of the degenerate star. Figure~\\ref{mass} shows the resulting masses corresponding to these two extreme K values. The masses are constrained by the fairly well-known inclination ($i\\sim70^{\\circ}-80^{\\circ}$) and by the requirement that the mass-losing star fills its Roche lobe. No main sequence star fills the Roche lobe defined by these plots, for orbital inclinations larger than $\\sim$30$^o$. Thus, the mass-losing star must have evolved off the main sequence to cause mass-exchange. The lowest mass that can thus evolve within a Hubble time is $\\sim$ 0.4 M$_{\\odot}$, and these are marked in the diagram at the appropriate inclination value. These therefore correspond to the lowest X-ray star masses under these assumptions, and just allow a white dwarf in the K=73 km s$^{-1}$ case. It may be possible that the mass-losing star has very low mass by having lost most of it in some earlier event, and is still filling its Roche lobe now. However, in any of these cases the X-ray star is more massive and the resulting masses are not those expected by the `standard' model (e.g. van den Heuvel et al.\\ 1992) in which it is assumed that the compact star is a $\\sim$1M$_{\\odot}$ white dwarf and the donor star has a mass of $\\sim$2.0 M$_{\\odot}$. The spectrum in eclipse shows no signs of a late type spectrum (see Figures~\\ref{med},~\\ref{binned},~\\ref{bin2}), consistent with the absence of an evolved more massive secondary. Unfortunately, the complex changes in the disk spectrum make it difficult to be more definitive than this at present, but it appears advisable to revisit the evolutionary scenarios to accomodate a mass-losing star that is less massive than the compact star, not only in CAL 87 but also for the other supersoft X-ray binaries. We discuss separately (Cowley et al.\\ 1998) the overall mass determinations for a number of supersoft binary systems. Some of the supersoft X-ray binaries have been found to have highly displaced lines indicating the presence of `jets'. These are most noticable in two low-inclination systems RX J0513$-$69 (Crampton et al.\\ 1996) and CAL 83 (Crampton et al.\\ 1987) where the displacements are several thousand kilometers per second. In the intermediate inclination system RX J0019$+$22 the jets show a velocity of $\\sim\\pm$800 km s$^{-1}$ from the central emission line. The jet lines move with the same phase and velocity amplitude as the central line, suggesting that the motion of He II 4686\\AA\\ emission is indeed orbital. Since CAL 87 is viewed nearly edge-on, any line emission coming from such jets would have a very low radial velocity and thus not be separated from the central emission profile. The asymmetry seen in the He II 4686\\AA\\ line of CAL 87 could arise from a pair of shifted lines which have larger amplitude (or possibly slightly different phase) than the central line. This would imply that the central line motion is not entirely orbital, but without better resolution this suggestion cannot be verified." }, "9804/astro-ph9804190_arXiv.txt": { "abstract": "We report on Westerbork 1.4 GHz radio observations of the radio counterpart to $\\gamma$-ray burst GRB~970508, between 0.80 and 138 days after this event. The 1.4 GHz light curve shows a transition from optically thick to thin emission between 39 and 54 days after the event. We derive the slope $p$ of the spectrum of injected electrons (d$N$/d$\\gamma_{\\rm e}\\propto\\gamma_{\\rm e}^{-p}$) in two independent ways which yield values very close to $p=2.2$. This is in agreement with a relativistic dynamically near-adiabatic blast wave model whose emission is dominated by synchrotron radiation and in which a significant fraction of the electrons cool fast. ", "introduction": "The peak luminosities of $\\gamma$-ray bursts (GRBs) are highly super-Eddington and require relativistic outflows (Paczy\\'{n}ski 1986; Goodman 1986). Paczy\\'{n}ski and Rhoads (1993) pointed out that radio emission is expected as a result of the interaction between such a relativistic outflow and an external medium, as is, e.g., observed in extragalactic jet sources (see also Katz 1994; M\\'{e}sz\\'{a}ros and Rees 1997). They estimated that the strongest GRBs may be followed by transient ($\\sim 10$ mJy) radio emission at intervals ranging from minutes (for a distance $d$ $\\sim$ 10$^{5}$ pc) to several weeks ($d$ $\\sim$ 10$^{9}$ pc). However, searches for radio counterparts through follow-up observations (Frail et al. 1994,1997a; Koranyi et al. 1995; Galama et al. 1997a,b) were without success until recently. With the rapid accurate location capability of the Wide Field Cameras (WFCs; Jager et al. 1995) onboard the Italian-Dutch X-ray observatory BeppoSAX (Piro et al. 1995) it has recently become possible to detect fading X-ray, optical and radio counterparts to GRBs (Costa et al. 1997a; Piro et al. 1997a; Groot et al. 1997a,b; Van Paradijs et al. 1997; Galama et al. 1997c,1998a; Sahu et al. 1997; Bond 1997; Metzger et al. 1997; Frail et al. 1997b,c; Bremer et al. 1998; Halpern et al. 1997). These observations have settled the discussion on the GRB distance scale (`galactic halo' versus `cosmological', see e.g. Fishman and Meegan 1995, Lamb 1995, Paczy\\'nski 1995): GRBs occur at Gpc distances. GRB 970508 is the first GRB to be detected in the radio (Frail et al. 1997b,c); the radio source position coincides with that of the optical (Bond 1997) and X-ray (Piro et al. 1997b) afterglow sources. Assuming that the variations of the source at 4.86 and 8.46 GHz are due to interstellar scintillation (ISS), their damping with time is consistent with a highly relativistically expanding shell passing a diameter of $\\approx 3\\mu$as (Frail et al. 1997b). VLBI observations show that the source is unresolved ($<$ 0.3 mas, Taylor et al. 1997). We here report on the results of 1.4 GHz radio observations of GRB 970508, made with the Westerbork Synthesis Radio Telescope (WSRT) between 0.80 and 138 days after the burst occurred. ", "conclusions": "} The observed optical spectral slope $\\alpha$ and the optical power law decay of the light curve $F_{\\nu} \\propto t^{\\delta}$ is not consistent with the expected relation for the simplest blast wave model ($\\delta = 3\\alpha/2$; e.g. Wijers, Rees and M\\'esz\\'aros 1997). The observed power law decay value, $\\delta = -1.141 \\pm 0.014$ ($t > 2$ days, Galama et al. 1998b; see also Pedersen et al. 1998, Castro-Tirado et al. 1998, Sokolov et al. 1998) would imply $\\alpha = -0.761 \\pm 0.009$, while in the optical passband $\\alpha = -1.12 \\pm 0.04$ is observed (Galama et al. 1998c, from here on Paper II; see also Sokolov et al. 1998). In the following we show that this may be explained by rapid cooling of a significant fraction of the electrons. A population of electrons with a power-law distribution of Lorentz factors $\\gamma_{\\rm e}$ (d$N$/d$\\gamma_{\\rm e}\\propto\\gamma_{\\rm e}^{-p}$) above some minimum value $\\gamma_{\\rm m}$ emits a power law synchrotron spectrum above the frequency $\\nu_{\\rm m}$ (corresponding to radiating electrons with $\\gamma_{\\rm m}$; e.g. Rybicki \\& Lightman 1979). Independently, above some Lorentz factor $\\gamma_{\\rm c}$ the electrons may cool rapidly, and an extra break in the spectrum is expected at the corresponding frequency $\\nu_{\\rm c}$ (Sari, Piran, \\& Narayan 1998). Beyond a certain time $t_0$ ($t_0$ is small $\\sim$ 500 sec; see Paper II) the evolution of the blast wave is adiabatic (Sari et al. 1998 and see Paper II); then $\\nu_{\\rm m} < \\nu_{\\rm c}$, and the spectrum varies as $F_{\\nu} \\propto \\nu^{-(p-1)/2}$ from $\\nu_{\\rm m}$ up to $\\nu_{\\rm c}$; above $\\nu_{\\rm c}$ it follows $F_{\\nu} \\propto \\nu^{-p/2}$ and below $\\nu_{\\rm m}$ it follows the low frequency tail, $F_{\\nu} \\propto \\nu^{1/3}$ (Sari et al. 1998). The evolution in time of the GRB afterglow is determined by the evolution of these break frequencies: $\\nu_{\\rm c} \\propto t^{-1/2}$ and $\\nu_{\\rm m} \\propto t^{-3/2}$ (both decrease with time). The decay part of the optical R$_{\\rm c}$ (Coussins R) band light curve (in the optical passband $\\nu > \\nu_{\\rm c}$ for $t\\gsim$ 1.2 days; Paper II) goes as $F_{\\nu} \\propto t^{(2-3p)/4}$, while the spectrum is then $F_{\\nu} \\propto \\nu^{-p/2}$ (Sari et al. 1998). This allows us to make two independent measurements of $p$: using $F_{\\rm R_{\\rm c}}\\propto t^{-1.141 \\pm 0.014}$ we find $p= 2.188 \\pm 0.019$ and using $\\alpha_{\\rm opt} = -1.12 \\pm 0.04$ gives $p = 2.24 \\pm 0.08$. The excellent agreement between the values of $p$ supports that a significant fraction of the electrons cool rapidly and that the evolution of the GRB remnant is adiabatic. Additional evidence for rapid cooling of a significant fraction of the electrons is given in Paper II. Observations by Bremer et al. (1998) with the IRAM Plateau de Bure Interferometer (PdBI) at 86 GHz show a maximum around $\\sim$ 12 days. We identify this maximum with the break frequency $\\nu_{\\rm m}$ passing 86 GHz at $t_{\\rm m,86 GHz} \\sim$ 12 days (Paper II). We expect, the 8.46 and 1.4~GHz emission to peak at $t_{\\rm m,8.46 GHz} \\sim$ 55 days and $t_{\\rm m,1.4 GHz} \\sim$ 180 days, respectively ($\\nu_{\\rm m} \\propto t^{-3/2}$). Near day 55, a shallow maximum can be seen in the 8.46 GHz light curve (Frail et al. 1997b). Unfortunately our 1.4 GHz light curve cannot be used to test the presence of the maximum at that frequency, both due to low signal to noise and because it ends 150 days after the burst, i.e. before the predicted maximum. Before $\\nu_{\\rm m}$ passes 8.46 GHz at $t_{\\rm m,8.46 GHz}$ we expect the 8.46 GHz spectrum to follow the low frequency tail $F_{\\nu} \\propto \\nu^{1/3}$, while after $t_{\\rm m,8.46 GHz}$ it is expected to be $F_{\\nu} \\propto \\nu^{-(p-1)/2} = \\nu^{-0.6}$ (Sari et al. 1998 and we have used $p$ = 2.2). Thus, we predict a gradual transition between $t_{\\rm m,8.46 GHz} \\sim$ 55 days and $t_{\\rm m,4.86 GHz} \\sim$ 80 days (when also at 4.86 GHz $\\nu_{\\rm m}$ has passed) from $\\alpha$ = 1/3 to $\\alpha = -0.6$. We note that this expectation is different from blast wave models that do not include the effect of rapid cooling of a significant fraction of the electrons ($F_{\\nu} \\propto \\nu^{-1.1}$ similar to the optical slope; see e.g. Wijers et al. 1997). Also the decays at 8.46 GHz (after $t_{\\rm m,8.46 GHz} \\sim$ 55 days) and 4.86 GHz (after $t_{\\rm m,4.86 GHz} \\sim$ 80 days) are expected to be different from that in the optical and X-ray passbands, $F_{\\nu} \\propto t^{3(1-p)/4} = t^{-0.9}$; where we have used $p$ = 2.2). These predictions can be tested with the continued monitoring at the VLA at 4.86 and 8.46 GHz by Frail et al. (1998). The radio afterglow light curves of GRB 970508 (Frail et al. 1997b and this Letter) show a much more gradual evolution than expected (see e.g. the fit to the 8.46 and 4.86 GHz data by Waxman, Kulkarni and Frail 1998). Also a constant self-absorption frequency was expected (e.g. Waxman et al. 1997) while we here show that a transition from optically thick to thin emission occurred around $\\sim$ 45 days. For $t < t_o$ Sari et al. (1998) predict a decrease with time of the self-absorption frequency, $\\nu_{\\rm a}$, while for $t >$ $t_0$ the self-aborption frequency $\\nu_{\\rm a}$ remains constant. The transition from optically thick to thin 1.4 GHz radiation then suggests that $t_0$ $\\sim$ 45 days. Also the 8.46 GHz light curve (Frail et al. 1997b) suggests that $t_0$ cannot be much smaller than 10 days, i.e. 10 days $\\lsim$ $t_0$ $\\lsim$ 55 days (we have extrapolated backwards in time from the 8.46 GHz peak at $t_{\\rm m}$ $\\sim$ 55 days with the expected dependence $F_{\\nu} \\propto t^{1/2}$ for times $t < t_{\\rm m}$). This is not in agreement with the finding that $t_0 \\sim 500$ sec (Paper II). However, the absence of a break in the smooth power law decay of the optical light curve from 2 to 60 days after the burst (Pedersen et al. 1998; Castro-Tirado et al. 1998; Sokolov et al. 1998; Galama et al. 1998c) shows that there is no important transition in that period. This does imply that some additional ingredient is needed; for example, Waxman et al. (1998) argue that the transition from ultrarelativistic to mildly relativistic expansion of the blast wave may explain the decrease in the self-absorption frequency $\\nu_{\\rm a}$ with time and the slow time dependence of the early radio light curves. The excellent agreement in the derived value for $p$ ($p$ = 2.2) from the decay of the optical light curve and the optical spectral slope support an adiabatic dynamical evolution of the GRB remnant and an extra break in the synchrotron spectrum at the frequency $\\nu_{\\rm c}$ above which the radiation is from electrons which cool rapidly compared to the remnant's expansion time. We predict a transition in the radio spectral index $\\alpha_{\\rm 4.86-8.46 GHz}$ from 1/3 to --0.6, between 55 and 80 days; the light curves are predicted to decay as $F_{\\nu} \\propto t^{-0.9}$ after 55 days at 8.46 GHz and 80 days at 4.86 GHz." }, "9804/astro-ph9804159_arXiv.txt": { "abstract": "We present observations of \\htwo\\ fluorescence at wavelengths between 1000 and 1200 \\AA\\ from the bright reflection nebula IC 63. Observations were performed with the Berkeley spectrograph aboard the \\orf-II mission \\cite{Hetal98}. To the best of our knowledge, this is the first detection of astrophysical \\htwo\\ fluorescent emission at these wavelengths (excluding planetary atmospheres). The shape of the spectrum is well described by the model of \\citeN{S89}. The absolute intensity, however, is fainter than an extrapolation from observations at longer ultraviolet wavelengths \\cite{WSBB89} by a factor of ten. Of the mechanisms that might help reconcile these observations, optical depth effects in the fluorescing \\htwo\\ itself are the most promising (or at least the most difficult to rule out). ", "introduction": "In many environments, the equilibrium abundance of the hydrogen molecule (\\htwo) is determined by a balance between formation on dust grains and photodissociation pumped by photons at wavelengths below about 1108 \\AA\\ \\cite{SB82}. Photons between this threshold energy and the Lyman limit populate a large number of rovibrational states in the $B^1 \\Sigma_u^+$ and $C^1 \\Pi_u$ electronic levels. En route back to the $X^1 \\Sigma_g^+$ level, the excited \\htwo\\ may radiate through any of several bound-bound and/or bound-free channels leading to a complex fluorescence emission spectrum. Located about 20$\\arcmin$ or 1.3 pc from $\\gamma$ Cas, dense gas in the reflection nebula IC 63 is illuminated by a bright UV radiation field. Line widths of a variety of molecular species are quite narrow \\cite{JVDB94}, suggesting that shock excitation is comparatively unimportant in this environment. The nebula provides an excellent test case for models of \\htwo\\ fluorescence and other photochemical processes \\cite{JVDBSS95}. The fluorescence model described in \\citeN{S88} and \\citeN{S89} has been successfully applied in the interpretation of infrared \\htwo\\ fluorescence from the reflection nebulae IC 63 and IC 59 \\cite{LLBJF97}, and to the ultraviolet fluorescence in the band near 1600 \\AA\\ from IC 63 \\cite{S89} and the Taurus molecular cloud \\cite{H94}. In this work we test the model in a previously unexplored band deep in the ultraviolet where the significant majority of the radiated fluorescent power is expected to emerge. ", "conclusions": "\\label{conclusions} We have detected \\htwo\\ fluorescence at wavelengths between 1000 and 1200 \\AA\\ from the bright reflection nebula IC 63 with the Berkeley spectrograph aboard the \\orf-II mission \\cite{Hetal98}. The wavelengths and relative strengths of the fluorescent features within the ORFEUS band agree well with the predictions of the model of \\citeN{S89}. The absolute fluorescent intensity is fainter than an extrapolation from observations at longer ultraviolet wavelengths \\cite{WSBB89} by a factor of ten. The measurements can not be reconciled by differential extinction from a foreground slab of dust (presumably associated with neutral gas) nor by absorption from quiescent \\htwo. Optical depth effects in the fluorescing \\htwo\\ itself, predicted by \\citeN{WSBB89}, remain the most plausible mechanism to explain our observations." }, "9804/astro-ph9804315_arXiv.txt": { "abstract": "We present contemporary infrared and optical spectra of the plateau type~II SN~1995V in NGC~1087 covering four epochs, approximately 22 to 84 days after shock breakout. The data show, for the first time, the {\\it infrared} spectroscopic evolution during the plateau phase of a typical type~II event. In the optical region P~Cygni lines of the Balmer series and of metals such as Sc~II, Fe~II, Sr~II, Ca~II and Ba~II lines were identified. The infrared (IR) spectra were largely dominated by the continuum, but P~Cygni Paschen lines and Brackett~$\\gamma$ lines were also clearly seen. The other prominent IR features are confined to wavelengths blueward of 11000~\\AA\\ and include Sr~II 10327, Fe~II 10547, C~I 10695 and He~I 10830~\\AA. Helium has never before been unambiguously identified in a type~IIp supernova spectrum during the plateau phase. We demonstrate the presence of He~I 10830~\\AA\\ on days 69 and 85. The presence of this line at such late times implies re-ionisation. A likely re-ionising mechanism is $\\gamma$-ray deposition following the radioactive decay of $^{56}$Ni. We examine this mechanism by constructing a spectral model for the He~I 10830~\\AA\\ line based on explosion model s15s7b2f of Weaver \\& Woosley (1993). We find that this does not generate the observed line owing to the confinement of the $^{56}$Ni to the central zones of the ejecta. In order to reproduce the He~I line, it was necessary to introduce additional upward mixing or ``dredge-up'' of the $^{56}$Ni, with $\\sim$10$^{-5}$ of the total nickel mass reaching above the helium photosphere. In addition, we argue that the He~I line-formation region is likely to have been in the form of pure helium clumps in the hydrogen envelope. The study of He~I 10830~\\AA\\ emission during the photospheric phase of core-collapse supernovae provides a promising tool for the constraint of initial mixing conditions in explosion models. ", "introduction": "Type~IIp (plateau) supernovae form the classic subgroup of the core-collapse supernovae. They are believed to arise from massive stars (12-25 M$_{\\odot}$) during the red supergiant phase. Early theoretical work by Falk and Arnett (1973) showed that hydrodynamical instabilities should appear in explosions of such massive stars. As the shock wave propagates through the stellar envelope it sets up density and pressure profiles which can in some cases result in the formation of Rayleigh-Taylor~(RT) instabilities (Chevalier 1976). In particular, RT instabilities are expected to grow at the interface of the core and the hydrogen envelope because of the large entropy (and density) jump that occurs there (Weaver \\& Woosley 1980). A direct consequence of these instabilities is that chemical mixing in the ejecta takes place (Bandiera 1984). Herant \\& Woosley (1994) have studied 2-D simulations of red supergiant explosions over a wide mass range, and found that the growth of hydrodynamic instabilities is highly likely in all cases. They showed that as the explosion (outgoing) shock plows into the hydrogen envelope, a reverse (ingoing) shock is formed, and between them, RT instabilities grow. Bubbles of hydrogen formed by these instabilities are violently dragged towards the centre of the star by the reverse shock. Simultaneously, compact helium and oxygen clumps advance out into the hydrogen envelope and bubbles of $^{56}$Ni are formed and distributed in the outer parts of the ejecta. It has also been realised that strong dredge-up should result from the neutrino-driven convection close to the neutron star surface (Herant \\& Benz 1992) which has been invoked to account for the conversion of the core-collapse to explosion. This should also produce fast-moving blobs or ``fingers'' of radioactive material which eventually penetrate the outer layers of the supernova. So far, only SN~1987A has provided us with clear observational evidence for dredge-up in a core-collapse event. This includes the shape of the light curve, the early detection of X-rays and $\\gamma$-rays and the width of the iron lines in the infrared. However SN~1987A was, of course, only a single event, and a rather unusual one in that it arose from a blue supergiant progenitor. Therefore, we cannot simply assume that similar dredge-up occurs in all other type~II events. Indeed, simulations have shown that differences in progenitor structure can lead to significantly modified hydrodynamical evolution (Herant \\& Benz 1992; Herant \\& Woosley 1994). Clearly, to establish whether or not deep dredge-up is typical of all core-collapse supernovae, a major step would be to demonstrate dredge-up and mixing in the most-common of all core-collapse events, the type~IIp supernova. A powerful demonstration of the occurrence of deep dredge-up would be the appearance of radioactive material at the surface at early times. Helium lines arising in the supernova envelope can be used as a tracer of the upwardly-mixed radioactive material. Helium lines are of high excitation. During the first week of the supernova, recombination maintains the populations of the excited levels and so He~I lines are seen. However after about 10 days, the conditions in the type~IIp atmosphere are such that all the helium will have recombined and de-excited to the ground state. But, if dredge-up occurs during the explosion, radioactive $^{56}$Ni may reach the outer parts of the supernova envelope at early times. If it does, the $\\gamma$-rays from its decay ($^{56}$Ni $\\Rightarrow$ $^{56}$Co $\\Rightarrow$ $^{56}$Fe) will excite or re-ionise the helium. Thus the detection of helium lines during the plateau phase (20-120~d post explosion) should imply upward mixing of radioactive material from the core. Unfortunately no optical He~I lines have ever been unambiguously identified during the plateau phase. However, there are two well-known strong lines in the infrared viz. He~I 10830~\\AA\\ (2s$^{3}$S--2p$^{3}$P$^{0}$) and 20580~\\AA\\ (2s$^{1}$S--2p$^{1}$P$^{0}$) which offer the prospect of testing for dredge-up of radioactive material. This technique was applied by Graham (1988) and Chugai (1991) using the 10830~\\AA\\ line in the SN 1987A at early times. Lucy (1991) invoked upward mixing of $^{56}$Ni to account for strong optical He I lines in the early-time spectra of type Ib supernovae. To investigate dredge-up in type~IIp supernovae we began in 1995 a programme of infrared and optical spectroscopy of this type of supernova. The data we present here comprise an extensive set of IR/optical spectroscopic observations of the type~IIp SN~1995V, spanning epochs of 22 to 85 days post explosion. The observations are described in section 2. In section 3 we compare the observations with a simple spectral synthesis model and discuss the line identifications, especially in the IR. In section 4 we describe the method we used to estimate the amount of dredge-up of $^{56}$Ni. In section 5 we present the results from the comparison of the model with the data, and our estimations for the amount of dredge-up. In section 6 we discuss the implications of this work for our understanding of dredge-up. ", "conclusions": "Reproduction by explosion model s15s7b2f of the observed He~I emission is achieved only by invoking substantial additional dredge-up of the $^{56}$Ni from the core. As expected, the case with no helium clumping requires the greatest dredge-up. However, we argue that the no-clumping case is probably highly unrealistic, since it is difficult to see how one could achieve such a large amount of $^{56}$Ni dredge-up and yet have no pure helium clumps in the same environment. We deduce, therefore, that there must exist some pure helium clumps in the hydrogen envelope. As we have shown, with fairly modest clumping of $\\chi_{He}=$0.1 or 0.2, the uniform central core coincides, to within the errors with that of the unmodified s15s7b2f model. The addition of a steep, power-law density component to this will bring sufficient $^{56}$Ni to the surface to account for the He~I 10830~\\AA\\ emission. Nevertheless, the fraction of the $^{56}$Ni mass which must be dredged up beyond the uniform core is quite substantial. For SN~1995V we conclude that a) a small amount of $^{56}$Ni ($\\sim$10$^{-6}$ M$_{\\odot}$) must have been dredged up to the helium photosphere (v$\\sim$4,250~km/s), and b) clumps of pure helium must have also existed in this region. High velocities for the decay products of $^{56}$Ni ($\\sim$ 3000 km/s) were also observed in the ejecta of SN 1987A (e.g. Meikle et al. 1993). As shown by Herant \\& Benz (1992), if the $^{56}$Ni is located at the base of the ejecta at t$\\sim$300s it is impossible to accelerate even a small fraction to about 3000 km/s during subsequent instabilities. In order to achieve such high velocities, it is necessary to invoke outward mixing of the nickel at even earlier times, such as might be caused by neutrino convection. If this occurred, then the later instabilities would carry the nickel to still higher velocities. In order to match the observations of SN1987A, Herant \\& Benz had to premix nickel out to 1.5 M$_{\\odot}$ above the mass cut. Herant \\& Woosley (1994) studied shock propagation, mixing and clumping in the explosion of red supergiants. In order to take into account the pre-mixing of $^{56}$Ni during the initiation of the explosion, they diluted the nickel by a factor $\\sim$4 above the mass cut. They then followed the shock propagation and the growth of RT instabilities. For all progenitors their simulations showed that extensive RT instabilities develop in the ejecta in the wake of the reverse shock from the H/He interface. In contrast to the blue supergiant studies, these instabilities have ample time in which to evolve and completely reshape the ejecta. In spite of this, in all the explosions simulated, nickel did not reach velocities higher than $\\sim$1500 km/s. Similarly, helium did not exceed velocities higher than 2500 km/s. Our results, therefore, indicate that a higher degree of pre-mixing may be required than is invoked in the Herant \\& Woosley models. Recently, Bazan \\& Arnett (1997) have simulated mixing in core-collapse events. Their simulations include both RT and Richtmeyer-Meshkov instabilities. These produce much higher velocities for $^{56}$Ni than do R-T instabilities alone. Velocities as high as $\\sim$4000~km/s are predicted. The mass and profile index of the upwardly-mixed $^{56}$Ni derived above could be of considerable value in constraining the initial parameters of these instabilities." }, "9804/astro-ph9804123_arXiv.txt": { "abstract": "s{ The X-ray emission from clusters of galaxies is one of the most pursued observational probe to investigate the distribution of dark matter and the related density parameter $\\Omega_0$. The crucial link to derive the statistics of observables from a dynamical theory is constituted by the physics for the diffuse baryons (or ICP) responsible of the X--ray emission. Here we present a physical model for the ICP which leads to a definite $L$--$T$ relation. Then we perform a physically based cosmological test, pointing out three cold dark matter universes: a Tilted critical CDM, a flat CDM with $\\Omega_0=0.3$, and an Open CDM with $\\Omega_0=0.5$, which are discussed on the basis of the RDCS survey. } ", "introduction": "Groups and clusters of galaxies constitute cosmic structures sufficiently close to equilibrium and with sufficient density contrast ($\\delta\\approx 2\\, 10^2$ inside the virial radius $R$) as to yield definite observables. They are dominated by dark matter (hereafter DM), while the baryon fraction is observed to be less than $20$\\%. The great majority of these baryons are in the form of {\\sl diffuse plasma} (ICP) with densities $n\\sim 10^{-3}$ cm$^{-3}$ and virial temperatures $k\\,T\\sim 5$ keV, and are responsible for powerful X--ray luminosities $L\\sim 10^{44}$ erg/s by optically thin thermal bremsstrahlung. As the plasma is a good tracer of the potential wells, much better than member galaxies, the X--ray emission is a powerful tool to investigate the mass distribution out to moderate and high redshifts. The ICP temperature directly probes the height of the potential well, with the baryons in the role of mere tracers; on the other hand, the luminosity with its strong dependence on density ($L\\propto n^2$) reliably probes the baryonic content and distribution. Statistically, an average $L$--$T$ correlation is observed along with substantial scatter, and this provides the crucial link to relate the X-ray luminosity functions with the underlying statistics of the DM. A physical model for the diffuse baryons is difficult to achieve. In fact, the simple self similar model (Kaiser 1986), which assumes the ICP amount to be proportional to the DM's at all $z$ and $M$, leads to a relation $L\\propto T^2$, conflicting with the observed correlation for rich clusters. The latter is close to $L\\propto T^{3.5}$ (David et al. 1993; Mushotzky \\& Scharf 1997). Here we propose a physical model for baryons, which leads to a prediction for the $L$--$T$ relation (see Cavaliere, Menci \\& Tozzi 1997, CMT97) and allows a non parametrical approach to the search for cosmological parameters. The results, presented in \\S 3, are a synthesis from Cavaliere, Menci \\& Tozzi (1998, CMT98). ", "conclusions": "We presented a physically based approach to cosmological tests with clusters of galaxies. We describe the X--ray emission from clusters with a specific model for the diffuse baryons. In this sense this approach is alternative to the parametrical approach (see Borgani this meeting). The results of our model depend on two parameters, the external temperature $T_1$ and density $n_1$, which are {\\sl not} free. Specifically, we use for $T_1$ the range $0.1\\div 0.8 $ keV provided by the stellar preheating. The value of $n_1$ for rich clusters is related to the DM density by the universal baryonic fraction. Thus we compute the expression of the bolometric luminosity for a given temperature. The average of the square of the density jump factor $\\langle g^2 \\rangle$ over the merging histories coupled with $\\beta(T)$ is what gives to the statistical $L-T$ correlation the curved shape shown in fig. \\ref{lt}b. In addition, our approach predicts an intrinsic {\\it variance} of dynamical origin due to the different merging histories, and built in the factor $g^2$. With the ICP state so described, we proceeded to constrain the cosmological parameters. After the observations by Rosati et al. (1998), we have computed the X-ray observables for groups and clusters of galaxies. On the basis of local data, the set of acceptable CDM universes is restricted to three disjoint domains: $\\Omega=1$ for the Tilted CDM with high baryon content; $\\Omega_o\\simeq 0.5$ for standard CDM; $\\Omega_o\\approx 0.3$ for CDM in flat geometry. However, only the TCDM and the $\\Lambda$CDM universes give acceptable faint counts. As an overall remark, a common feature of all the above universes is constituted by some excess in the counts. This may indicate some non--trivial incompleteness in the canonical hierarchical clustering, worth keeping under scrutiny. We recall that in the adiabatic models for the ICP (Evrard \\& Henry 1991, Kaiser 1991) the evolution of the $L-T$ relation is reduced or even negative, thus alleviating the excess. However, the anti--evolution required in OCDM would be very difficult to justify (the adiabatic models are largely discussed in CMT98). Now the question is: to what extent enlarging the data base on X--ray clusters will help in further constraining cosmology? We argue that the variance intrinsic to the hierarchical clustering, and amplified by the ICP emissivity, sets an effective limitation. Richer, faint surveys will hardly provide a sharper insight into cosmology unless one reduces both the uncertainty concerning $\\sigma_8$ and the larger one concerning $L_{o}$. However, we stress that such efforts will find soon a more proper aim than constraining $\\Omega_o$. This is because MAP, and subsequently PLANCK, will accurately measure on still linear scales not only the perturbation power spectrum but also directly $\\Omega_o$. Once the cosmological framework has been fixed, the study of groups and clusters in X-rays will resume its proper course, that is, the physics of systems of intermediate complexity which is comprised of the DM and of the ICP component." }, "9804/astro-ph9804065_arXiv.txt": { "abstract": "We present an empirical method which measures the distance to a Type Ia supernova (SN~Ia) with a precision of $\\sim$ 10\\% from a single night's data. This method measures the supernova's age and luminosity/light-curve parameter from a spectrum, and the extinction and distance from an apparent magnitude and color. We are able to verify the precision of this method from error propagation calculations, Monte Carlo simulations of well-sampled SNe~Ia, and the Hubble diagram of scarcely observed supernovae. With the reduction in telescope time needed, this method is three to four times more efficient for measuring cosmological parameters than conventional light-curve based distance estimates. ", "introduction": "The explosion of a Type Ia supernova (SN~Ia) is a catastrophic phenomenon veiled in layers of complexity. Recent efforts to monitor these events have led to an increased ability to predict, if not fully understand, the stages of SN Ia evolution. The model for the photometric history of SNe~Ia has been refined from a homogeneous description \\cite[]{brunophd,branch_miller,ts95} to one which characterizes a relation between peak luminosity and light-curve shape \\cite[]{philm15,hametal95,hametal96,rpk95,kim_stretch98}. The slower, broader light-curves are intrinsically brighter at peak than the faster, narrower light-curves. Recognizing and exploiting such relations has led to a renaissance in the use of SNe~Ia as extragalactic distance indicators. Extending luminosity/light-curve relations to multiple passbands separates the competing effects of dust, intrinsic differences, and distance on the light of SNe~Ia \\cite[]{rpk96}. Distances with 5-10\\% uncertainty can be obtained using the light-curve shapes of well-observed supernovae. The optical spectra of SNe~Ia are rich in information [see \\cite{fil97} for a review]. Many of the elements synthesized and ejected in the explosion have been identified despite the blending of their high-velocity profiles \\cite[]{bran81b,nugphd}. In addition, the relative strengths of some spectral features have been shown to correlate with SN~Ia peak luminosity \\cite[]{nugseq95}. As the supernova evolves, predictable casts of features appear and disappear, illuminated by the photosphere's recession through the synthesized layers. The temporal evolution of these features is sufficiently reliable to be used as a clock to determine the current age of a SN~Ia to a precision of 1-2 days \\cite[]{mink39,riess_age97}. Unfortunately, supernovae occur without warning, making it difficult to collect the observations necessary to measure their distances. Observing an unscheduled event in up to four filters many times over the course of $\\sim$ 100 days is a time consuming and logistically formidable task. The observing record for a typical SN~Ia is quite fragmentary. Following this process, it will be many years of work to gather the number of SN~Ia distances necessary to put strong limits on cosmological parameters. Even at high redshifts ($z \\geq 0.3$), where a strategy for batch detections of multiple supernovae has made it possible to schedule supernova discoveries and their follow up \\cite[]{perl97}, difficulties arise. Since these observations require the largest telescopes, the light-curves are typically more poorly sampled than the nearby ones caught at a similar phase. Yet, from a single night's observations, a SN~Ia's spectrum and photometric magnitudes can reveal its age, intrinsic luminosity, extinction, and apparent brightness. From this information one can estimate the distance to a supernova without further observations (except for the possible need of a galaxy image to subtract the host's light). Here we explore this possibility with two independent sets of SNe~Ia. We describe this technique in \\S 2 and its expected uncertainty in \\S 3. In \\S 4 we apply it to randomly selected snapshots of extensively observed SNe~Ia to determine the precision of such distance estimates. In \\S 5 we construct the Hubble diagram of ``cast-off'' SNe~Ia: objects which were observed only once or a few times. We extend the application of this method in \\S 6 to SNe~Ia with $0.2 \\leq z \\leq 0.83$. In \\S 7 we discuss variations of this technique and its leverage on estimating cosmological parameters. ", "conclusions": "In principle, enough information can be garnered from a single supernova spectrum and photometric epoch to estimate the distance to a SN~Ia. In practice, the results of \\S 3, 4, and 5 suggest this method produces distances having a precision of $\\sim$ 10\\%, with variations that are a function of the quality of the data and the age of the supernova. Depending on the amount of host galaxy contamination, it may be necessary to obtain spectra and images of the host galaxy after the supernova has faded. The snapshot distance method employs the same luminosity and extinction corrections used in the MLCS method of Riess, Press, and Kirshner 1996a (and more recently updated in Riess et al. 1998b). We find no significant offset between the distance estimates of the two methods and only a modest reduction in precision for the snapshot distance method. There are three more limited versions of a distance method using single epoch SN Ia observations which reveal the utility of luminosity and extinction corrections (see Table~\\ref{discomp_tab} and Figure~\\ref{show}). These variants employ the SFA measurement but lack the luminosity correction, the extinction correction, or both. Disregarding individual luminosities and light-curve shapes predicted by the spectral ratios, we fit homogeneous, fiducial templates to the photometric epoch. Further, we discarded our estimate of the extinction from the color excess. We measured the resulting ``standard candle'' distances to SNe~Ia using the Monte Carlo technique described in \\S 4. As seen in Figure~\\ref{show}, the distribution of dispersions has a mean of 0.35 mag, a value consistent with previous SNe~Ia distance estimates which assume SN Ia light curve homogeneity and do not correct for extinction \\cite[]{st93,ts95,hametal95,hametal96,rpk95,rpk96,branch_miller,vauetal95}. These distances are also 15\\% {\\it greater} in the mean (or smaller in the implied Hubble constant) than either MLCS or snapshot distances, consistent with other comparisons of distance estimates which assume homogeneity instead of heterogeneity of SNe~Ia \\cite[]{ts95,hametal95,hametal96,rpk95,rpk96}. To simulate the effect of single filter information, we used our luminosity correction without an extinction correction. This procedure results in a mean dispersion of 0.25 mag and distances which are only 3\\% greater than those obtained from MLCS. This result, though better than the standard candle method, is still worse than the complete snapshot procedure. A final variant is to disregard a luminosity/light-curve shape correction but maintain an extinction correction from the color excess. Such a method using light-curves was proposed by \\cite{vdb95} to account for both intrinsic luminosity differences as well as absorption by dust. This method takes advantage of the coincidental near-agreement between the standard reddening law and the relation between intrinsic color and luminosity to correct for both extinction and luminosity differences. \\cite{rpk96} have noted that while both sources of luminosity variation affect the SN color in the same direction, the specific ratios of the luminosity difference to color difference are not precisely the same for extinction and intrinsic SN Ia variation. Monte Carlo simulations of this method combined with a SFA measurement give a mean dispersion of 0.19 mag and distances which are 5\\% greater in the mean than those of the MLCS method. Despite the low dispersion of this method, we are suspicious of the distances it predicts. The distribution of dispersions obtained from our Monte Carlo simulation (see Figure~\\ref{show} and Table~\\ref{discomp_tab}) is more skewed than any other, including an asymmetric ``tail'' encompassing dispersions greater than 0.3 mag. We believe that for SNe~Ia with only moderate amounts of extinction or whose luminosities are similar to those of typical SNe~Ia, this method has merit. Yet for very red SNe~Ia, this method can predict distances which are systematically and considerably in error due to the inability to distinguish between absorption by dust and intrinsic variation. The snapshot method predicts distances which agree in the mean with only a moderate reduction in precision from light-curve shape methods. Yet because of the greatly diminished expense in data collection, this method can be more effective for problems which benefit equally well from a high {\\it quantity} of SNe~Ia distances as from the {\\it quality} of those distances. Two such applications are mapping the nearby peculiar velocity field and determining the cosmological parameters which dictate global geometry. Recent attempts to map the cosmic velocity field with SNe~Ia \\cite[]{riess_beta97} suffer from dilute spatial sampling. Replete peculiar velocity maps could reveal the influence of matter fluctuations and constrain the matter content of the local Universe. Nearby, many more SNe~Ia are discovered than can be regularly monitored. By decreasing the observational requirements of each SN Ia, it should be possible to increase the sampling of the local velocity field. The light-curves employed by \\cite{hametal96} and \\cite{rpk96} are typically sampled for 10 to 15 epochs. The observational demands increase as the supernova rapidly fades. With the telescope time invested in a single set of SN Ia multi-color light-curves, sufficient data for 15 to 25 SNe~Ia snapshot distances could be gathered. Accounting for the inherent distance uncertainties, telescope time spent collecting snapshot distance data is 3 to 4 times more efficient than time spent collecting light-curves for distance estimates. Efforts to measure the cosmological parameters $\\Omega_M$ and $\\Omega_\\Lambda$ from distant SNe~Ia could also profit from snapshot distances. Systematic searches for SNe~Ia at z $\\geq 0.3$ have yielded a plethora of objects \\cite[]{perliau95a,schetal95}. Combining a new generation of large telescopes with $\\sim$ 1 degree fields of view and multi-fiber spectrometers with the snapshot method could allow SNe~Ia distances to be gathered in batch at an unprecedented rate. At the current rate of discovery, a night spent searching five 1-degree fields followed by a night collecting spectra of the candidates with a multi-fiber spectrometer could yield $\\sim$ 50 SNe~Ia distances \\cite[]{schmidt97,rate_96}. Repeating this process every new moon could yield up to $\\sim$ 600 distances a year. At this rate of accumulation, it should be possible to convincingly separate the effects of various sources of energy density on the redshift-magnitude relation \\cite[]{omol_95}. A more optimal method for measuring SN Ia distances would employ both the predictive power of SN Ia light and color curve shapes with that of SN Ia spectra. Such a method would replace the distinction between a snapshot distance and a light-curve distance with a distance estimate which makes the most economical use of all available SN Ia observations. Using the tools described in Riess, Press, and Kirshner 1996a, Nugent et al. 1995, Riess et al. 1997, Riess et al. 1998b, and this paper, such a method appears to be quite feasible. \\bigskip This work was supported by the NSF through grant AST--9417213 to A.V.F. and AST-9528899 and AST-9617058 to R.P.K., by the Miller Institute for Basic Research in Science through a fellowship to A.G.R., and by the Director, Office of Computational and Technology Research, Division of Mathematical, Information, and Computational Sciences of the U.S. DoE under contract number 76SF00098 to P.E.N. Some of the calculations presented in this paper were performed at the National Energy Research Supercomputer Center (NERSC), supported by the U.S. DoE. We thank Stephan Benetti and George Djorgovski for allowing us to use their spectra of SNe Ia prior to publication, and Bruno Leibundgut for suggestions that helped improve this paper. \\appendix" }, "9804/hep-ph9804285_arXiv.txt": { "abstract": "The expected proton and neutrino fluxes from decays of massive metastable relic particles are calculated using the HERWIG QCD event generator. The predicted proton spectrum can account for the observed flux of extremely high energy cosmic rays beyond the Greisen-Zatsepin-Kuzmin cutoff, for a decaying particle mass of ${\\cal O}(10^{12})$~GeV. The lifetime required is of ${\\cal O}(10^{20})$~yr if such particles constitute all of the dark matter (with a proportionally shorter lifetime for a smaller contribution). Such values are plausible if the metastable particles are hadron-like bound states from the hidden sector of supersymmetry breaking which decay through non-renormalizable interactions. The expected ratio of the proton to neutrino flux is given as a diagonistic of the decaying particle model for the forthcoming Pierre Auger Project. ", "introduction": "It has been known for some time that interactions on the 2.73 K blackbody cosmic microwave background (CMB) will severely degrade the energies of cosmic ray nucleons with energies beyond $\\sim5\\times10^{19}\\ev$ --- the Greisen-Zatsepin-Kuzmin (GZK) cutoff \\cite{gzk}. It was therefore very surprising when the Fly's Eye atmospheric fluorescence detector reported the observation of an extremely high energy cosmic ray (EHECR) event with an energy of $(3.0\\pm0.9)\\times10^{20}\\ev$ \\cite{flyseye}. This was followed by the detection of a $(1.7-2.6)\\times10^{20}\\ev$ event by the AGASA air shower array \\cite{agasa}. These discoveries substantiated earlier claims from the Volcano Ranch \\cite{vr}, Haverah Park \\cite{hp} and Yakutsk \\cite{yak} air shower arrays that cosmic rays do exist beyond the GZK cutoff. About a dozen such events are now known. Detailed accounts of the data may be found in recent reviews \\cite{reviews}. In Figure~\\ref{fig1} we show the EHECR spectrum for energies exceeding $10^{18}\\ev$ \\cite{spec}; note that the fluxes have been multiplied by $E^3$. It is believed that cosmic rays with energies up to $\\sim5\\times10^{18}\\ev$, the so-called `ankle', are predominantly of galactic origin, possibly accelerated by the Fermi mechanism in supernova remnants \\cite{books}. Above this energy, the spectrum flattens and the composition changes from being mostly heavy nuclei to mostly protons. Such a correlated change in the spectrum and composition was first established by the Fly's Eye experiment \\cite{flyseye} and Figure~\\ref{fig1} shows their suggested two-component fit to the data. The new component which dominates at energies beyond $\\sim5\\times10^{18}\\ev$ is isotropic and therefore cannot possibly originate in the galactic disk \\cite{agasa2,lc95}. However it also extends well beyond the GZK cutoff raising serious problems for hypothetical extragalactic sources. Because of the rapid energy degradation at these energies through photo-pion production on the CMB, such sources must exist within $\\sim500\\mpc$, in fact within $\\sim50\\mpc$ for the highest energy Fly's Eye event \\cite{cronin}. For heavy nuclei, the energy loss is less severe according to a revised calculation \\cite{heavy} so the range may extend upto $\\sim100\\mpc$. General arguments \\cite{greisen,hillas} provide correlated constraints on the magnetic field strength and spatial extent of the region necessary to accelerate particles to such high energies and these requirements are barely met by likely astrophysical sites such as active galactic nuclei and the `hot spots' of radio galaxies \\cite{cr}. Moreover there are few such sources close to us and no definite correlations have been found between their locations and the arrival directions of the most energetic events \\cite{source,agasa2}. It has been speculated that gamma-ray bursts which too are isotropically distributed, may be responsible for EHECRs \\cite{grb}. However since these are at cosmological distances, one would expect to see the GZK cutoff in the cosmic ray spectrum contrary to observations (cf. ref.\\cite{grbtest}). Some of the above arguments may be evaded if the EHECR events are due not to nucleons but neutral particles such as photons and neutrinos. Although high energy photons also suffer energy losses in traversing the CMB and the extragalactic radio background, there is no threshold effect which would cause a cutoff near the GZK value \\cite{photon}. However the observed shower profile of the highest energy Fly's Eye event \\cite{flyseye} argues against the primary being a photon since it would have interacted on the geomagnetic field and started cascading well before entering the atmosphere \\cite{notphoton}. The observed events are also unlikely to be initiated by neutrinos as they all have incident angles of less than $40^\\circ$ from the zenith and thus too small a path length in the atmosphere for interactions \\cite{gqrs96}. This argument may be evaded if neutrinos become strongly interacting at high energies due to new physics beyond the Standard Model \\cite{neutrino,hongmo}, but such proposals are found not to be phenomenologically viable \\cite{bhg98} (although this is disputed \\cite{nu}). (Alternatively, the propagating high energy neutrinos could annihilate on the relic cosmic neutrino background, assumed to have a small mass of ${\\cal O}(0.1)$~eV, to make hadronic jets within the GZK zone \\cite{weiler}.) Other exotic possibilities have been suggested, e.g. monopoles \\cite{wk96}, stable supersymmetric hadrons \\cite{farrar} and loops of superconducting cosmic string (`vortons') \\cite{bp97}. However these possibilities have many phenomenological problems \\cite{mn98,ber} and we do not discuss them further. Thus one is encouraged to seek `top-down' explanations for EHECRs in which they originate from the decay of massive particles, rather than being accelerated up from low energies. The most discussed models in this connection are based on the annihilation or collapse of topological defects such as cosmic strings or monopoles formed in the early universe \\cite{hill,td,bs95,bv97}. When topological defects are destroyed their energy is released as massive gauge and Higgs bosons which are expected to have masses of ${\\cal O}(10^{16})\\gev$ if such defects have formed at a GUT-symmetry breaking phase transition. The decays of such particles can generate cascades of high energy nucleons, $\\gamma$-rays and neutrinos. A more recent suggestion is that EHECRs arise from the decays of metastable particles with masses $m_X\\sim10^{13}-10^{16}\\gev$ which constitute a fraction of the dark matter \\cite{bkv97}. These authors suggest that such particles can be produced during reheating following inflation or through the decay of hybrid topological defects such as monopoles connected by strings, or walls bounded by strings. The required metastability of the particle is ensured by an unspecified discrete symmetry which is violated by quantum gravity (wormhole) effects. Another suggestion is that the long lifetime is due to non-perturbative instanton effects \\cite{kr97}. In ref.\\cite{fkn97}, a candidate metastable particle is identified in a $SU(15)$ GUT. A generic feature of these `top-down' models is that the EHECR spectrum resulting from the decay cascade is essentially determined by particle physics considerations. Of course the subsequent propagation effects have astrophysical uncertainties but since the decays must occur relatively locally in order to evade the GZK cutoff \\cite{bkv97}, they are relatively unimportant. Thus although the proposal is speculative, it is possible, in principle, to make reliable calculations to confront with data. In this work we consider another possible candidate for a relic metastable massive particle \\cite{ben98} whose decays can give rise to the observed highest energy cosmic rays. First we discuss (\\S~\\ref{crypton}) why this candidate, which arises from the hidden sector of supersymmetry breaking, is perhaps physically better motivated than the other suggested relics. We then undertake (\\S~\\ref{decay}) a detailed calculation of the decay cascade using a Monte Carlo event generator to simulate non-perturbative QCD effects. This allows us to obtain a more reliable estimate of the cosmic ray spectrum than has been possible in earlier work on both topological defect models \\cite{td} and a decaying particle model \\cite{bkv97}. We confront our results with observations and identify the mass and abundance/lifetime required to fit the data. We conclude (\\S~\\ref{concl}) with a summary of experimental tests of the decaying particle hypothesis. ", "conclusions": "We have investigated the hypothesis that the highest energy cosmic rays, in particular those observed beyond the GZK cutoff, arise from the decay of massive metastable relic particles which constitute a fraction of the dark matter in the galactic halo. To simplify computations (using the HERWIG Monte Carlo event generator) we have considered only decays into $q\\bar{q}$ pairs with unit branching ratio. Comparison with experimental data indicates that a decaying particle mass of ${\\cal O}(10^{12})\\gev$ is required to fit the spectral shape while the absolute flux requires a lifetime of ${\\cal O}(10^{20})\\yr$ if such particles contribute the critical density. The predicted decay spectra may be somewhat altered if 3-body decays and other final states (e.g. supersymmetric particles \\cite{bk97}) are considered. However our conclusions regarding the preferred mass and relic abundance/lifetime of the decaying particle are unlikely to be affected. In particular it would appear that the approximations used to calculate the particle spectra in previous studies of decaying topological defects \\cite{td} and hypothetical massive particles \\cite{bkv97} were not sufficiently accurate. Our work indicates that the topological defect model is disfavoured unless the mass of the decaying gauge bosons is less than about $10^{13}\\gev$, which is well below the unification scale of $\\sim10^{16}\\gev$. (A similar conclusion is arrived at by independent arguments in refs.\\cite{bss97,vah98}.) By contrast, cryptons from the hidden sector of supersymmetry breaking have a mass of the required order, as well as a decay lifetime which is naturally suppressed. However their relic abundance is difficult to estimate reliably, although we have argued that it may be cosmologically interesting. The primary intention of this work is to attempt to quantify the decaying particle hypothesis in a manner which is of interest to experimentalists. We have therefore computed the expected neutrino to proton ratio as a function of energy since this is an important test of competing hypotheses for forthcoming experiments, in particular the Pierre Auger project \\cite{auger}. Of course our cleanest prediction is that the cosmic ray spectrum should cut off just below the mass of the decaying crypton, at $\\sim3\\times10^{20}\\ev$. Moreover, with sufficient event statistics it should be possible to identify the small anisotropy which should result from the distribution of the decaying particles in the Galactic halo \\cite{dt98}. Thus although the hypothesis investigated here is very speculative, it is nevertheless testable. Perhaps Nature has indeed been kind to us and provided a spectacular cosmic signature of physics well beyond the Standard Model." }, "9804/astro-ph9804175_arXiv.txt": { "abstract": "s{We are contructing an interferometric telescope, the Very Small Array, to study the cosmic microwave background on angular scales 0.2--$4.5^{\\circ}$. The physical layout and electronic design of the telescope are optimised to give maximum protection from systematic effects, while still providing sufficient sensitivity to make high signal-to-noise images. A prototype single baseline is currently being tested, with scientific results expected during 2000.} ", "introduction": "It is widely accepted that the majority of the cosmological results that might be obtained from the cosmic microwave background (CMB) will come from accurate measurements of the CMB power spectrum over the region of the accoustic peaks, that is in the range $100 < l < 2000$ (where $l$ is the spherical harmonic multipole). We have therefore been constructing an instrument, the Very Small Array (VSA), to measure the power spectrum (and provide images) in precisely this angular range. We have argued previously\\cite{jones} that interferometers are very well suited for ground-based CMB measurements, given their relative immunity to the atmosphere and other systematic effects compared with switched-beam experiments. Here we will describe some of the more detailed design considerations of the array, and review the progress of the project so far. ", "conclusions": "" }, "9804/astro-ph9804205_arXiv.txt": { "abstract": "We present HST WFPC2 observations in three bands (F450W=B, F467M and F814W=I) of a group of three galaxies at $z=2.8$ discovered in a ground-based narrow-band search for \\lya emission near the $z=2.8$ quasar \\pks. One of the galaxies is a damped \\lya (DLA) absorber and these observations bear on the relation between the DLA clouds and the Lyman-break galaxies and the stage in the evolution of galaxies they represent. We describe a procedure for combining the undersampled WFPC2 images pointed on a sub-pixel grid, which largely recovers the full sampling of the WFPC2 point spread function (psf). These three galaxies have similar properties to the Lyman-break galaxies except that they have strong \\lya emission. The three galaxies are detected in all three bands, with average $m_B\\sim26$, $m_I\\sim25$. Two of the galaxies are compact with intrinsic (i.e. after correcting for the effect of the psf) half-light radii of $\\sim0.1$ arcsec ($0.4 h^{-1}$ kpc, $q_{\\circ}=0.5$). The third galaxy comprises two similarly compact components separated by 0.3 arcsec. The HST images and a new ground-based \\lya image of the field provide evidence that the three galaxies are more extended in the light of \\lya than in the continuum. Combined with the evidence from the \\lya line widths, previously measured, this suggests that we are measuring the size of the surface of last scattering of the escaping resonantly-scattered \\lya photons. The measured impact parameters for this DLA galaxy (1.17 arcsec), for a second confirmed system, and for several candidates, provide a preliminary estimate of the cross-section-weighted mean radius of the DLA gas clouds at $z\\sim 3$ of $<13 h^{-1}$ kpc, for $q_{\\circ}=0.5$. The true value is likely substantially smaller than this limit as DLA clouds at small impact parameter are harder to detect. Given the observed sky covering factor of the absorbers this implies that for $q_{\\circ}=0.5$ the space density of DLA clouds at these redshifts is more than five times the space density of spiral galaxies locally, with the actual ratio probably considerably greater. For $q_{\\circ}=0.0$ there is no evidence as yet that DLA clouds are more common than spiral galaxies locally. We summarise evidence that filamentary structures occur in the distribution of galaxies at high redshift. ", "introduction": "By studying typical galaxies at high redshift, here $z>2$, we can record the origins of normal galaxies such as our own Milky Way. Schmidt \\cite{sc65} was the first to observe a high$-$redshift galaxy when he obtained a spectrum of 3C9, a radio$-$loud quasar of redshift $z=2.01$. Normal star$-$forming high$-$redshift galaxies are about 1000 times fainter, and remained undiscovered until the present decade. A small number of candidates (some of which may be active galaxies) have been identified with 4$-$metre telescopes, through the detection of Ly$\\alpha$ emission (e.g. Steidel, Sargent, and Dickinson 1991, Lowenthal et al. 1991, M\\o ller and Warren 1993, Pascarelle et al. 1996, Francis et al. 1995). Another approach has been to employ deep broad$-$band imaging to identify candidates by the expected Lyman$-$limit discontinuity in their spectra (Steidel and Hamilton 1992). The brightest Lyman$-$break galaxies have $m_R\\sim 24$ and spectroscopic confirmation has only become feasible with the completion of the Keck 10$-$metre telescope. The report by Steidel et al. (1996) on the results of spectroscopic observations of Lyman$-$break candidates really marked the beginning of the statistical study of high-redshift galaxies. They were able to confirm redshifts for 15 star$-$forming galaxies in the range $3.02$. Complementary to the searches for starlight from high$-$redshift galaxies have been the analyses of absorption lines in the spectra of high$-$redshift quasars. These studies have provided measurements of the mass density of neutral hydrogen in the universe (e.g. Wolfe et al. 1986, Lanzetta et al. 1991), and the metallicity of the gas (e.g. Pettini et al 1994), and how these quantities have changed with redshift. The analysis of Pei and Fall (1995) of the absorption--line data reconstructs the global history of star formation, gas consumption, and chemical enrichment, accounting in a self--consistent way for the effects of the progressive extinction of the background quasars due to dust as star formation proceeds. The advantage of the absorption--line approach to the history of star formation is that it is global in nature as all the neutral gas at any redshift is directly observed. With the deep imaging studies it is necessary to extrapolate below the survey flux limit to measure the total star formation rate. The weakness of the global approach however is that it tells us nothing about the sizes and morphologies of the galaxies in which the star formation occurs. Deep imaging of DLA absorbers provides the missing information. In other words if we imaged the absorbers we could combine the data of faint galaxy surveys (luminosities, sizes, shapes) with the information on the gas obtained from the spectroscopic studies (column densities, chemistry), allowing a more detailed comparison with theories of galaxy formation. Unlike for MgII absorbers (e.g. Bergeron and Boiss\\'e 1991, Steidel, Dickinson, and Persson 1994a) little progress has been made in such a programme of imaging of DLA systems. There are only two published unambiguous detections, discussed below, as well as a small number of candidate counterparts of DLA absorbers (at both low redshift $-$ Steidel et al. 1994b, Le Brun et al. 1997 $-$ and high-redshift $-$ Arag\\'{o}n-Salamanca et al. 1996). In this paper we report on the results of 30 orbits of imaging observations with the {\\it Hubble Space Telescope} (HST) of a group of three galaxies at $z=2.81$, named S1, S2, S3, one of which (S1) is a damped \\lya absorber. The three galaxies lie in the field of the quasar \\pks, and were detected in the light of \\lya emission by M\\o ller and Warren (1993, hereafter Paper I). A detailed discussion of their nature was presented by Warren and M\\o ller (1996, hereafter Paper II \\footnote{We note here that recent spectroscopic observations by Ge et al. (1997) confirm our conclusion that the \\lya emission from the DLA absorber is due to star formation rather than photoionisation by the quasar.}). The HST images provide measures of their sizes, magnitudes, and colours. The other high$-$redshift damped system that has been successfully imaged is the absorber at $z=3.150$, towards the quasar $2231+131$, observed by Djorgovski et al. \\cite{dj96}. In this paper we intercompare the measured properties of these two absorbers, the two companion galaxies S2 and S3, and the population of Lyman$-$break galaxies, to draw conclusions about the space density of DLA absorbers, their structure, and the relation between the DLA absorbers and the Lyman-break galaxies. The layout of the rest of the paper is as follows. The HST observations of the field towards \\pks are described in Section 2, as well as new ground$-$based narrow$-$band observations of the field. The three galaxies are very small, and the HST images were dithered with a half-integer pixel step in order to improve the image sampling. The algorithm for combining the images is outlined in this section. In Section 3 we present the reduced HST images. In Section 4 we provide the results of aperture photometry and profile fitting of the HST and ground-based images. Finally, in Section 5 we discuss these results and their implications for our understanding of the nature of DLA absorbers and Lyman-break galaxies. ", "conclusions": "\\subsection{Continuum emission} The three galaxies S1, S2, S3 are similar in their properties to members of the recently detected population of Lyman-break galaxies, except that they have strong \\lya emission, with mean restframe ${\\rm EW=60\\AA}$. For example the B and I magnitudes for our three sources $m_B\\sim 26$, $m_I\\sim 25$, are within the range of magnitudes of the Lyman-break galaxies of redshift $z\\sim3$ observed by Steidel et al. (1996) and Lowenthal et al. (1997). The average half-light radius of the four objects listed in Table 4 is 0.1 arcsec. This is smaller than the value quoted for the Lyman-break galaxies (Giavalisco et al. 1996, Lowenthal et al. 1997), by a factor of two to three. However we have broken S3 into two sub components and measured the radii for each subcomponent. In addition the radii quoted in Table 4 have been corrected for the effect of the psf. Young star-forming galaxies should have approximately constant flux per unit frequency over the restframe ultraviolet and optical regions of the spectrum. The expected colour of a source at $z=2.8$ with a power-law continuum varying as $f_{\\nu}\\propto\\nu^0$, after accounting for typical absorption in the \\lya forest, is $m_B-m_I=0.66$. As seen in Table 3 the $m_B-m_I$ colours for the three sources S1, S2, S3 are consistent with this value, for both the small and large apertures. The fact that S1 has properties similar to the Lyman-break galaxies suggests that DLA absorbers and Lyman-break galaxies may commonly be associated with each other. Lowenthal et al (1997) have also made a connection between the two populations by suggesting that the broad \\lya absorption line seen in several cases in the spectra of Lyman-break galaxies is a damped line from neutral hydrogen in the galaxy. However the absorption lines seen in their average spectrum and in the relatively high S/N spectrum presented by Ebbels et al. (1996) are not optically thick in the line centre. The absorption--line profiles are difficult to interpret because the background source is extended. It is even possible that in some cases the absorption line is stellar absorption in the spectra of late B stars and that the galaxies are observed in a post-burst phase, rather than actively forming stars (see Valls-Gabaud 1993) i.e. for some cases this may be the explanation for the weakness of the \\lya emission line in the Lyman-break galaxies, as opposed to dust. High-redshift galaxies with strong \\lya emission, like the three sources near \\pks, may therefore be younger than many of the Lyman-break galaxies. \\subsection{Extended \\lya emission} In this subsection we summarise the evidence that the \\lya emission from the three sources is more extended than the continuum emission. The evidence for S2 and S3 comes from an examination of the $m_B-m_M$ and $m_B-m_N$ colours, summarised in Table 3. The bottom row in Table 3 provides the predicted $m_B-m_M$ colours for each source computed, as detailed in Section 4.1.2, from the B and N total magnitudes. The small aperture $m_B-m_M$ colours are smaller than the predicted values, as would be the case if the \\lya emission is more extended than the continuum. For S2 the difference is significant at $2.8\\sigma$, and for S3 at $2.3\\sigma$. Notice also that the large aperture $m_B-m_M$ colours are larger than the small aperture colours, which is consistent with the hypothesis of extended \\lya emission. A broad absorption line in the M passband could only explain part of the discrepancy between the measured and predicted colours. For example for S2, simply removing $50{\\rm\\AA}$ of the continuum flux within the M band changes the $m_B-m_M$ colour by only 0.1 mag. The total B magnitude for S1 (i.e. the large-aperture measurement from the HST image, Table 4) is not reliable because of the uncertainty associated with the psf subtraction. This prevents us from usefully making the same comparison between predicted and measured $m_B-m_M$ colours for S1. However there is an indication that S1 may also be more extended in \\lya. The measurements of the half-light radius of S1 from the ground-based \\lya image yielded values of $0.51^{+0.08}_{-0.08}$ arcsec for the exponential profile and $0.66^{+0.25}_{-0.19}$ arcsec for the de Vaucouleurs profile (Section 4.2.2). These values are significantly larger than the half light radius of the region of continuum emission of $0.13\\pm0.06$ arcsec (Table 4), measured from the HST image. Such a diffuse contribution of \\lya to the M-band flux would be hard to detect, and could have been partially subtracted in attempting to remove large-scale residuals from the psf subtraction (Section 4.2.1). However we cannot rule out the possibility that there is also a similar diffuse contribution to the continuum flux. The argument for extended \\lya is made visually in Figure 4. Here we compare the M-band images after subtraction of the computed contribution of the continuum (i.e. showing the \\lya emission only), against how the objects would have appeared if the \\lya light profile were the same as the continuum profile. \\begin{figure} \\vspace{8.5cm} \\special{psfile=Reffig_1.ps angle=0 voffset=-52 hoffset=-14 vscale=40 hscale=40} \\caption[ ]{Illustration of the evidence for extended \\lya emission for, from top to bottom, S1, S2, S3. Each panel shows the same 2 arcsec by 2 arcsec region as Figure 3, here smoothed by a 0.15 by 0.15 arcsec boxcar filter to enhance faint features. The left column shows the summed image from all three filters. The middle column shows the M-band images with the computed contribution of the continuum subtracted. This image, then, contains only the \\lya line component. In the right column we show how the \\lya-only image would have appeared if the \\lya profile were identical to the continuum profile.} \\end{figure} To summarise, there is evidence that the regions of \\lya emission from these three sources are more extended than the regions of continuum emission. Although the colour and size differences quoted above are only marginally significant, these results accord with our earlier conclusion (Paper II) that the cause of the relatively broad \\lya emission lines for S1 and S3 is resonant scattering, as the photons escape though a high column density of HI. In this picture the \\lya image records the last scattering surface, which will be larger than the region of star formation. We note in passing that this implies that the observed strength of \\lya absorption and emission in spectra can depend on the slit width used. \\subsection{The sizes of the DLA gas clouds} In this section we make a summary of the measured impact parameters $b$ of DLA absorbers. Using this we estimate the cross-section-weighted mean radius $\\bar{R}_{DLA}$ of the gas clouds at $z>2$. Combining this information with the measured line density of DLA absorbers $dn/dz$ we are able to infer the ratio of the comoving space density of DLA absorbers at $z>2$ to the local space density of spiral galaxies. \\begin{table*} \\begin{flushleft} \\caption[]{Measured impact parameters $b$ of DLA absorbers} \\begin{tabular}{lcccccc} \\hline\\noalign{\\smallskip} Quasar&$z_{abs}$&log(N(HI))&$b$&$b$& confirmed & reference \\\\ & & &($q_{\\circ}=0.0$)&($q_{\\circ}=0.5$)& & \\\\ & &cm$^{-2}$ &kpc&kpc& & \\\\ \\hline $0454+0393$ & 0.86 & 20.8 & $4.1h^{-1}$& $3.3h^{-1}$ & N & 1 \\\\ $0302-223$ & 1.01 &$\\leq20.0\\:\\:\\:\\:\\:$& $6.2h^{-1}$& $4.9h^{-1}$ & N & 1 \\\\ $1331+170$ & 1.78 & 21.2 & $4.7h^{-1}$& $3.1h^{-1}$ & N & 1 \\\\ $0151+048$A & 1.93 & 20.4 & $7.7h^{-1}$& $5.0h^{-1}$ & Y & 2 \\\\ $1215+333$ & 2.00 & 21.0 & $8.4h^{-1}$& $5.3h^{-1}$ & N & 3 \\\\ $0841+129^a$& 2.37 & 21.3 & $8.0h^{-1}$& $4.7h^{-1}$ & N & 3,4 \\\\ $0841+129^a$& 2.48 & 21.0 & $8.0h^{-1}$& $4.7h^{-1}$ & N & 3,4 \\\\ $0528-250^b$& 2.81 & 21.3 & $8.1h^{-1}$& $4.5h^{-1}$ & Y & 5 \\\\ $2231+131$ & 3.15 & 20.0&$15.7h^{-1}\\:\\,$& $8.2h^{-1}$ & Y & 5 \\\\ \\noalign{\\smallskip} \\hline \\end{tabular} \\\\ $^a$ the detected galaxy could be the counterpart to one or other of the two DLA absorbers listed \\\\ $^b$ impact parameter taken from HST image (this paper) \\\\ References: 1. Le Brun et al. (1997), 2. Fynbo et al. (1997), 3. Arag\\'{o}n-Salamanca et al. (1996), 4. Wolfe et al. (1995), 5. M\\o ller and Warren (1993), 6. Djorgovski et al. (1996). \\end{flushleft} \\end{table*} There have been many searches for optical counterparts of the DLA absorbers. In Table 6 we have listed the small number of DLA absorbers, with redshifts $z>0.8$, for which likely candidate or confirmed optical counterparts have been detected. Listed are the quasar name, the absorber redshift and column density, the measured impact parameter (i.e. the projected physical separation between the optical counterpart and the line of sight to the quasar), and whether or not the counterpart has been confirmed, i.e. the redshift of the counterpart has been measured to be the same as the redshift of the absorber. In Figure 6 we plot impact parameter against column density for the absorbers listed in Table 6, for $q_{\\circ}=0.5$. The detection of the counterpart to a DLA absorber by broad-band imaging, which requires the digital subtraction of the quasar image, is more difficult for small impact parameters where the photon noise from the quasar image is greater. Therefore it is very likely that the measured impact parameters of the few absorbers for which counterparts have been detected are larger than the average impact parameter for DLA absorbers. The vertical line in Figure 6 at ${\\rm log(N(HI))=20.3}$ marks the conventional lower limit of the column density of DLA absorbers. The distribution of points in the figure is consistent with the expected anticorrelation between impact parameter and column density. We are interested in the mean impact parameter $\\bar{b}_{DLA}$ that would be measured if counterparts were detected for all DLA absorbers in a large unbiased survey. From inspection of Figure 6 we suggest that it is safe to conclude that for $z>2$, $q_{\\circ}=0.5$, $\\bar{b}_{DLA}$ is less than $7h^{-1}$ kpc. For $q_{\\circ}=0.0$ we suggest $\\bar{b}_{DLA}<13h^{-1}$ kpc. The actual mean impact parameters are probably substantially smaller than these limits. We relate the mean impact parameter to the cross-section-weighted mean radius $\\bar{R}_{DLA}$ of the absorbers by $\\bar{b}_{DLA}=\\alpha\\bar{R}_{DLA}$. The edge of the DLA is the point at which the column density falls below $N(HI)=2\\times10^{20}$ cm$^{-2}$, and we effectively assume that the column density falls off sharply at larger radii. For face on disks $\\alpha=\\frac{2}{3}$, and for a disk that is nearly edge on $\\alpha=\\frac{4}{3\\pi}$, so we take $\\alpha=0.55$ as an average value for randomly inclined disks. On this basis we infer the following limits on the cross-section-weighted mean radius of DLA absorbers at high redshift: $$\\bar{R}_{DLA}<23.6h^{-1} {\\rm kpc}\\:\\:\\:\\: (z>2, q_{\\circ}=0.0) $$ $$\\bar{R}_{DLA}<12.7h^{-1} {\\rm kpc}\\:\\:\\:\\: (z>2, q_{\\circ}=0.5). $$ Using these limits and the measured line density of absorbers $dn/dz$ we can compute limits to the ratio of the comoving space density of DLA absorbers to the the local space density of spiral galaxies. We follow essentially the methodology employed by Wolfe et al. (1986), and Lanzetta et al. (1991, hereafter LTW). For spiral galaxies locally they adopted a galaxy luminosity function $\\Phi(\\frac{L}{L_*})$ of Schechter form, with power-law index $s$, a power-law (Holmberg) relation between radius and luminosity, of index $t$, and a ratio between gas radius and optical (Holmberg) radius $\\xi$, independent of luminosity. They found that the incidence of DLA absorbers per unit redshift $dn/dz$ at $z\\sim 2.5$ was considerably higher than expected, by a factor $F\\sim 5$, on the basis of no evolution in galaxy cross section or luminosity function normalisation. We now allow for evolution by supposing that the space density of DLA absorbers is higher than the local space density of spirals by a factor $E_{\\Phi}(z)$, and that the gas radii of galaxies are larger at high redshift by a factor $E_r(z)$. Since $dn/dz$ is proportional to the product of the space density and the galaxy cross section $\\sigma$, we have that $F=E_{\\Phi}(z)E^2_r(z)$. By comparing the expected value of the cross-section-weighted radius $\\bar{R}$ to the measured limits to $\\bar{R}_{DLA}$, we obtain limits to $E_r(z)$. Then from the measured values of $F$ we determine limits to $E_{\\Phi}(z)$, which is our goal. Under the above assumptions the cross-section-weighted average radius of DLA absorbers is given by: $$\\bar{R}(z)=\\frac{\\int_0^{\\infty} R(\\frac{L}{L_*})\\sigma(\\frac{L}{L_*}) \\Phi(\\frac{L}{L_*})d(\\frac{L}{L_*})} {\\int_0^{\\infty}\\sigma(\\frac{L}{L_*})\\Phi(\\frac{L}{L_*})d(\\frac{L}{L_*})}$$ $$=\\frac{E_r(z)\\xi R_*\\Gamma(1+3t-s)}{\\Gamma(1+2t-s)}$$ where $R_*$ is the optical radius of a local $L_*$ spiral galaxy. Following LTW we adopt the following values of the parameters: $t=0.4$, $s=1.25$, $\\xi=1.5$, $R_*=11.5h^{-1}$ kpc. This leads to $\\bar{R}(z)=11.0h^{-1}E_r(z)$. Comparing with the above measured limits to $\\bar{R}_{DLA}$ we obtain the following limits to the growth factor of galaxy disks: $$E_r(z)<2.15\\:\\: (z>2, q_{\\circ}=0.0)$$ $$E_r(z)<1.16\\:\\: (z>2, q_{\\circ}=0.5)$$ For their sample D2, of which at least 30 out of the 38 candidate DLA absorbers have been confirmed, LTW found $F=3.8$, for $q_{\\circ}=0.0$, and $F=7.1$, for $q_{\\circ}=0.5$. These results then imply that the ratio of the comoving space density of DLA absorbers at high redshift to the local space density of spiral galaxies is given by: $$E_{\\Phi}(z)=\\Phi_*(DLA)/\\Phi_*(spiral)>0.8\\:\\: (z>2, q_{\\circ}=0.0) $$ $$E_{\\Phi}(z)=\\Phi_*(DLA)/\\Phi_*(spiral)>5\\:\\: (z>2, q_{\\circ}=0.5) $$ Because the actual average impact parameters of DLA absorbers are very likely substantially smaller than the quoted limits, the actual comoving space density ratios are probably considerably greater than the limits quoted above. \\begin{figure} \\vspace{11.5cm} \\special{psfile=Fig_bvsN.ps angle=0 voffset=-60 hoffset=-35 vscale=50 hscale=50} \\caption[ ]{Relation between impact parameter and column density for all confirmed or candidate counterparts of DLA absorbers of redshift $z>0.8$ (listed in Table 6). Different symbols correspond to different redshift ranges: squares $z>2.5$; triangles $2.5\\geq z>1.5$; circles $1.5\\geq z>0.8$. Absorbers for which the counterpart has been confirmed by spectroscopy are shown as filled symbols, and are labeled. The two possible DLA absorbers for the quasar $0841+129$ are joined by a dotted line. The hatched area shows the densely populated region from the simulation of Katz et al. (1996), and the dashed lines are the boundaries enclosing all the points in the simulation.} \\end{figure} In Figure 5 we show also the relation between impact parameter and column density measured by Katz et al (1996) from a hydrodynamic simulation of a CDM universe with $q_{\\circ}=0.5$. Given the limited spatial resolution of the simulation, and the fact that star formation and consequent feedback were not treated, it would be premature to draw any conclusions about the apparent good agreement between the results of the simulation and the observations. Nevertheless this plot and the above calculation underscore the importance of measuring impact parameters for a large sample of DLA absorbers, a goal we are pursuing by imaging with the STIS instrument on HST. \\subsection{The structure of the DLA absorbers} The observations summarised above provide some indication of the typical structure of a DLA absorber. The following sketch is suggested. At the centre of the gas cloud, corresponding in projection to the highest column densities, there may be a region of star formation, of diameter $\\sim 1h^{-1}$ kpc. This central source would be observed as a Lyman-break galaxy. The gas column density decreases outwards, as suggested by the anticorrelation between impact parameter and $N_{HI}$. The size of the region over which the column density is $>2\\times10^{20}$cm$^{-2}$ is several kpc. The region of star formation would be surrounded by a zone of ionised gas. The \\lya photons are resonantly scattered in escaping to the surface of the surrounding cloud of neutral gas. The size of the observed region of \\lya emission is larger than the region of star formation, but smaller than the diameter of the gas cloud. The latter indicates a preferred direction of escape, implying that the gas resides in a flattened structure. In reality the gas cloud is likely to be irregular rather than smooth, and might contain several knots of star formation, surrounded by HII regions of different sizes, and the \\lya photons will escape preferentially where the column density of neutral gas is lowest, possibly along a complex network of tunnels. Merging clouds will be observed as galaxies with irregular structure. \\subsection{Filamentary structure in the distribution of galaxies at $z=3$} We have previously suggested (Paper II) that the approximately linear arrangement of S1, S2, S3, as well as other known groups at high-redshift, may correspond to the filamentary arrangements of galaxies and galaxy sub-units that are found in computer simulations of the high-redshift universe. Figure 6 summarises the observational situation, and shows the spatial arrangement of four high-redshift groups of galaxies discovered in \\lya searches (from top to bottom: Francis et al., 1995, considering the \\lya absorber as a galaxy; Pascarelle et al., 1996; Paper I; Le F\\`{e}vre et al., 1996). All groups have been rotated to a horizontal baseline. Clearly each of the structures is elongated. \\begin{figure} \\vspace{9.5cm} \\special{psfile=Fig_align.ps angle=0 voffset=-145 hoffset=-78 vscale=58 hscale=58} \\caption[ ]{Relative positions of galaxies in four high-redshift groups discovered in \\lya searches. Each group has been rotated so that the principal axis is horizontal in the figure, and the groups have been offset from one another vertically. Note the different scale of the lower plot.} \\end{figure} To quantify the degree of alignment of each group we have measured the smallest internal angle $\\delta_L$ of the triangle defined by the three objects in each group (for the PKS 0528-250 field at z=2.81 we used the position of S1 for the DLA absorber, rather than the position of the quasar sightline through the absorber). The measured angles are, from top to bottom in the figure, $\\delta_L = 7.0^{\\circ}, 1.0^{\\circ}, 5.8^{\\circ}, 2.1^{\\circ}$. The average value of the angle $\\delta_L = 4.0^{\\circ}$ is certainly very suggestive of filamentariness in the distribution of galaxies at high redshift. It is also notable that the sizes of the groups range over an order of magnitude. In the computer simulations filaments are seen at all scales (e.g. Evrard et al. 1994). We do not suggest that every high-redshift group of three or four galaxies discovered will exhibit a similar degree of alignment, but rather that filamentary networks akin to those seen in the simulations may become visible as larger samples of high-redshift galaxies with denser sampling become available. The phenomenon might provide a useful discriminant for models of structure formation." }, "9804/astro-ph9804033_arXiv.txt": { "abstract": "Recent observations supported by theoretical models have lead to the view that giant and supergiant stars are over abundant, and/or a high metallicity component may be present, in the stellar populations at the centres of active galaxies. Here we attempt to quantify these effects by observing the strengths of the stellar absorption lines of Mg~b, NaI, CaII triplet as well as molecular bands such as CN and TiO. Using long-slit spectroscopic data we are able to separate the stellar populations in and around the nucleus, for a sample including, normal, LINER, starburst and Seyfert galaxies. In this paper we present the data, namely spectra of the nucleus and of a number of circum-nuclear regions. Comparisons reveal gradients in both the reddening and the stellar population within the central regions of most galaxies. Detailed stellar population synthesis is presented in a companion paper. ", "introduction": "A crucial unsolved question is whether the stellar populations in the nuclear regions of Active Galaxies differ from those of non-active galaxies of the same Hubble type. Correlations between near IR CO indices, far-infrared and X-ray luminosities of active galaxies may indicate that the more powerful monsters reside in more actively star-forming host galaxies (Yamada, 1994). The accelerated star formation caused by dynamic instabilities which trigger and/or fuel the nuclear activity could result in an overabundance of giant, supergiant and super metal rich (SMR) stars (Scoville, 1992 and references therein). Terlevich et al. (1990) observed that in some Active Galactic Nuclei (AGN) the near IR CaII triplet absorption features are as strong, or even stronger than those of normal non-active galactic nuclei. They have suggested that the ``featureless\" blue continuum previously thought to be non-stellar in origin, actually arises from the unresolved continuum from a young cluster of stars containing red supergiants. The presence of a ``truly'' featureless continuum is then required only in the case of some Seyfert~1 galaxies which show dilution of the CaII triplet lines. Based on independent evidence we know that in some galaxies a starburst region surrounds the unresolved nucleus which emits the broad lines e.g. NGC~7469 (Wilson et al., 1991). Nonetheless this does not prove a causal connection between the starburst and AGN. An alternative interpretation, plausible within the framework of our knowledge of AGN, is that the stellar lines are superimposed on a non-stellar nuclear continuum, and are formed in a super metal rich population. Indeed, one could expect abundance anomalies due to the intense star formation in the metal rich environment of the nucleus. From our detailed study of the Seyfert~1 nucleus in the galaxy NGC~3516 (Serote Roos et al., 1996), we find that the stellar population exhibits a noticeable dilution by a featureless continuum in the wavelength range 5000-9800\\AA\\ {\\it as well as} a high proportion of super metal rich stars. Thus, before dismissing the possibility of the presence of any non-thermal component in the near infrared continuum of Seyfert galaxies, we must first define the stellar population in the nucleus and the surrounding regions for a sample of AGN of all levels of activity. Until now most studies tackling this subject have made use of only a small spectral domain around a few spectral features. In order to make further progress it is necessary to extend these studies to cover more stellar absorption features, providing signatures of different stellar populations e.g. MgI $\\lambda$5175, NaI~D $\\lambda$5896 and 8196\\AA\\ and the numerous TiO and CN bands. The NaI lines, for example, help to distinguish between the effects of overmetallicity rather than a supergiant dominated population. Indeed, the ambiguity between metallicity and supergiant excess can be resolved by comparison of the CaII triplet and NaI absorption strengths (e.g. Zhou, 1991) although the interstellar contribution to NaI introduces some uncertainty, and also by comparison with the strength of MgI (Couture \\& Hardy, 1990). In this paper we present long-slit spectroscopy, in the range 5000-9800\\AA\\, of a sample of galaxies with different levels of activity. These observations are used to estimate radial gradients in the stellar population and to extract the stellar spectrum of the very nucleus. The sample is selected to include various classes of active galaxies, i.e. Seyferts 1 and 2, LINERs and starbursts, in order that the strength of the stellar features may be correlated with the general properties of each class. We also present data for two normal galaxies which will be used as comparison templates. This is not a complete sample in any statistical sense, but it does provide insights into the diversity of bulge stellar populations in and around AGN of different classes of activity. The detailed analysis of the stellar populations for this sample using a new spectral synthesis programme (Pelat, 1997), is to be found in Paper II (Serote Roos et al., 1997). ", "conclusions": "The nuclei of our sample galaxies are generally redder than the outer regions (this excludes Seyfert 1s as their nuclear spectra are dominated by strong broad emission lines). This colour gradient could either be due to dust or to stellar population gradients. In addition to dust/population gradients, the presence of a featureless continuum is inferred in a number of AGN. This component might be of non-thermal origin, plausible in Seyfert galaxies, or of stellar origin in the case of nuclear starburst activity. The extra component dilutes the strengths of the stellar absorption lines in the spectra of the nuclear regions of AGN. A detailed study of the spectral shape of the diluting component will allow us to differentiate between the two hypotheses. A full population synthesis analysis for most of the galaxies in our sample, using a new method developed to determine unique solutions, is presented in Paper II." }, "9804/astro-ph9804080_arXiv.txt": { "abstract": " ", "introduction": "Ascertaining the core collapse supernova mechanism is a long-standing problem in astrophysics. The current paradigm begins with the collapse of a massive star's iron core and the generation of an outwardly propagating shock wave that results from core rebound. Because of nuclear dissociation and neutrino losses, the shock stagnates. This sets the stage for a shock reheating mechanism whereby neutrino energy deposition via electron neutrino and antineutrino absorption on nucleons behind the shock reenergizes it (Bethe \\& Wilson 1985; Wilson 1985). The shock reheating phase is essential to the supernova's success, but it is precisely this phase that is difficult to simulate realistically. During shock reheating, core electron neutrinos and antineutrinos are radiated from their respective neutrinospheres, and a small fraction of this radiated energy is absorbed in the exterior shocked mantle. The shock reheating depends sensitively on the electron neutrino and antineutrino luminosities, spectra (best characterized by the {\\small RMS} energies), and angular distributions in the region behind the shock (e.g., see Burrows \\& Goshy 1993, Janka \\& M\\\"{u}ller 1996, Mezzacappa \\etal 1998). These, in turn, depend on the neutrino transport in the semitransparent region encompassing the neutrinospheres, necessitating a neutrino transport treatment that is able to transit accurately and seamlessly between neutrino-thick and neutrino-thin regions. Various neutrino transport approximations have been implemented in simulating core collapse supernovae. The most sophisticated approximation, which naturally has been used in realistic one-dimensional simulations, is multigroup flux-limited diffusion ({\\small MGFLD}; e.g., Bowers \\& Wilson 1982, Bruenn 1985, Myra \\etal 1987). {\\small MGFLD} closes the neutrino radiation hydrodynamics hierarchy of equations at the level of the first moment (the neutrino flux) by imposing a relationship between the flux and the gradient of the neutrino energy density (the zeroth moment). For example, \\begin{equation} F_{\\nu}=-\\frac{c\\Lambda}{3}\\frac{\\partial U_{\\nu}}{\\partial r}+..., \\label{eq:mgfld} \\end{equation} \\begin{equation} \\Lambda = \\frac{1}{1/\\lambda + |\\partial U_{\\nu}/\\partial r|/3U_{\\nu}}, \\label{eq:lambda} \\end{equation} \\smallskip \\noindent where $\\lambda$ is the neutrino mean free path, and $U_{\\nu}$ and $F_{\\nu}$ are the neutrino energy density and flux (Bruenn 1985). [Other forms for the flux-limiter $\\Lambda$ can be found in Bowers \\& Wilson (1982), Levermore \\& Pomraning (1981), and Myra \\etal (1987).] Whereas the limits $\\lambda \\rightarrow 0$ and $\\lambda \\rightarrow \\infty$ produce the correct diffusion and free streaming fluxes, it is in the critical intermediate regime where the {\\small MGFLD} approximation is of unknown accuracy. Unfortunately, the quantities central to the postshock neutrino heating, i.e., the neutrino {\\small RMS} energies, luminosities, and mean inverse flux factors, are determined in this regime, and given the sensitivity of the neutrino heating to these quantities, it becomes necessary to consider more accurate transport schemes. Moreover, in detailed one-dimensional simulations that have implemented elaborate {\\small MGFLD} neutrino transport (e.g., see Bruenn 1993, Wilson \\& Mayle 1993, and Swesty \\& Lattimer 1994), explosions were not obtained unless the neutrino heating was boosted by additional phenomena, such as convection. This leaves us with at least two possibilities to consider: (1) Failures to produce explosions in the absence of additional phenomena, such as convection, have resulted from neutrino transport approximation. (2) Additional phenomena may be essential in obtaining explosions. ", "conclusions": "Comparing three-flavor {\\small MGBT} and three-flavor {\\small MGFLD} in postbounce supernova environments, we find that {\\small MGBT} leads to a significant increase/decrease in the {\\it net} heating/cooling rate, particularly just above/below the gain radius. The {\\small MGBT} net heating rate can be as much as 2 times the {\\small MGFLD} net heating rate above the gain radius, with net cooling rates that are typically 0.8 times the {\\small MGFLD} rate below. These differences stem primarily from differences in the neutrino luminosities and mean inverse flux factors; the heating rate is linearly proportional to both these quantities, and differences in both add to produce a significant difference in the net heating rate. We also observe that the differences in the net heating rate are greatest at earlier postbounce times for a given progenitor mass, and at a given postbounce time, greater for greater progenitor mass. This is illustrated in Table 1. The enhancement in heating with increased progenitor mass suggests that the net heating enhancement from {\\small MGBT} is potentially robust and self-regulated. In closing, our results are promising, and their ramifications for core collapse supernovae, and in particular, for the postbounce neutrino-heating, shock-revival mechanism, await one- and two-dimensional dynamical simulations with {\\small MGBT} coupled to the core hydrodynamics. One-dimensional simulations are currently underway, and we plan to report on them soon. \\begin{table}[t] \\vspace{-24pt} \\begin{center} \\footnotesize\\rm \\caption{Maximum Net Heating/Cooling Rates \\label{t1}} \\begin{tabular}{lccc} \\topline {Progenitor Mass [${\\rm M}_{\\odot}$]} & {$t_{pb}$ [ms]} & {Maximum Net Heating Ratio} & {Maximum Net Cooling Ratio} \\\\ \\midline 15& 106& 2.0& 0.8 \\\\ & 233& 1.3& 0.8 \\\\ 25 & 156& 2.0& 0.8 \\\\ \\bottomline \\end{tabular} \\end{center} \\vspace{-12pt} \\end{table}" }, "9804/astro-ph9804049_arXiv.txt": { "abstract": "We investigate in detail the hypothesis that low surface brightness galaxies (LSB) differ from ordinary galaxies simply because they form in halos with large spin parameters. We compute star formation rates using the Schmidt law, assuming the same gas infall dependence on surface density as used in models of the Milky Way. We build stellar population models, predicting colours, spectra, and chemical abundances. We compare our predictions with observed values of metallicity and colours for LSB galaxies and find excellent agreement with {\\it all} observables. In particular, integrated colours, colour gradients, surface brightness and metallicity match very well to the observed values of LSBs for models with ages larger than 7 Gyr and high values ($\\lambda > 0.05$) for the spin parameter of the halos. We also compute the global star formation rate (SFR) in the Universe due to LSBs and show that it has a flatter evolution with redshift than the corresponding SFR for normal discs. We furthermore compare the evolution in redshift of $[Zn/H]$ for our models to those observed in Damped Lyman $\\alpha$ systems by \\scite{Pettini+97} and show that Damped Lyman $\\alpha$ systems abundances are consistent with the predicted abundances at different radii for LSBs. Finally, we show how the required late redshift of collapse of the halo may constrain the power spectrum of fluctuations. ", "introduction": "Late type low surface brightness galaxies are found to be a significant fraction of the total galaxy population in the Virgo cluster (\\pcite{Impey+88}), in the Fornax cluster (\\pcite{Irwin+90}), and in the field (\\pcite{Mcgaugh+95}; \\pcite{Sprayberry+96}; \\pcite{Sprayberry+97}). A theory of galaxy formation should therefore account for the existence of LSBs. On the other hand, LSBs are particularly useful because they are simpler objects than high surface brightness (HSB) galaxies: they have relatively little present day star formation and little dust (reddening). Therefore it is easier to model their stellar population and star formation history. They are also more dark matter dominated than HSBs, and therefore particularly suitable for probing the structure of dark matter halos. In order to compare theories of galaxy formation with observational surveys of galaxies it is important not only to quantify observationally the abundance of LSBs, but also to understand their nature. For instance, if LSBs are made of very young stellar populations, as often claimed in the literature, they may be irrelevant in the picture of the Universe at high redshift, while they could be a substantial component of that picture if their stellar populations are old. The existence of LSB galaxies can be readily understood if it is assumed, following \\scite{Fall+80}, that the specific angular momentum of the baryons is approximately conserved during their dissipation into a rotationally supported disk, and that the disk length--scale is therefore related to the angular momentum of the dark matter halo. The low surface brightness is the consequence of the low surface density of the disk, which is due to the larger spin parameter of the dark matter halos of LSBs relative to the spin parameter of the halos of HSB disks (e.g. \\pcite{Dalcanton_disc+97}; \\pcite{Mo+97}; \\pcite{Jimenez+Heavens+97}). However, while the surface density of a disk can be simply related to the mass and spin parameter of its dark matter halo, its surface brightness, and therefore its mass to light ratio ($M/L$), depends on the stellar population and could in principle vary with galaxy type and age. For instance, $M/L$ must have a significant time dependence due to the history of star formation in the disk and to the continuous death of massive stars. Moreover, $M/L$ in a given photometric band changes with time also as a consequence of the colour evolution of the stellar population, which is sensitive also to the history of the chemical enrichment of the disk. Therefore, a comparison between a simple model of disk formation and the observations cannot be made without an appropriate model for the stellar population in the disk that includes {\\it self-consistently} the chemical evolution of the population. \\begin{figure} \\centering \\leavevmode \\epsfxsize=1. \\columnwidth \\epsfbox{surf_1p0.eps} \\caption[]{Initial surface density for an isothermal sphere and 3 different masses. The solid lines correspond to $\\lambda=0.1$ to 0.01 (from top to bottom) in steps of 0.01. The circular velocity of the halo (dashed line) is also plotted.} \\label{f1} \\end{figure} In this paper we model self-consistently the chemical and photometric evolution of the stellar population of LSB disks, formed in dark matter halos described by isothermal spheres. The aim of this paper is twofold: \\begin{enumerate} \\item show that only one parameter, namely the {\\it spin parameter $\\lambda$} of the dark halo, can explain the surface brightness of LSBs. \\item use LSB disk galaxies to study the formation and evolution of all disk galaxies. \\end{enumerate} The second point is motivated by the first, that is by the fact that LSB disks are in fact so similar to HSB disks. On the other hand, LSBs are more dark matter dominated than HSBs, and therefore better described by a very simple model where only the gravitational field of an isothermal halo is considered. Our assumptions are as follows: \\begin{enumerate} \\item{The specific angular momentum of baryonic matter is the same as for dark matter} \\item{Gas settles until centrifugally supported, in a given dark matter halo} \\item{The star formation rate is given by the Schmidt law} \\item{The gas infall rate is assumed to be the same function of total surface density as used in models of the Milky Way} \\end{enumerate} \\begin{figure*} \\centering \\leavevmode \\epsfxsize=1.6 \\columnwidth \\epsfbox{surf_o.eps} \\caption[]{Contour plot of the value of the central surface density (in M$_{\\odot}$pc$^{-2}$) for different values of the halo mass and spin parameter. Left panel: halo formation redshift $z=0.7$. Right panel: halo formation redshift $z=2.1$.} \\label{f2} \\end{figure*} We first use the density profile of the halo to compute the surface density of the settling disk. We then use the Schmidt formation law and an infall rate that reproduces the observed properties of the Galaxy in conjunction with the chemical evolution models by \\scite{Matteucci+89} to compute the star formation rate at several radii of the disk and the evolution of several chemical species (H, D, He, C, N, O, Ne, Mg, Si, S, Ca, Fe and Zn). We then proceed to compute spectra, integrated colours, colour profiles and surface brightness for two different values of the spin parameter of the halo. We first present in the next section the simple non--self--gravitational disk model and the non--singular halo model. We describe in section 3 the set of synthetic stellar population models used to predict the spectra and colours of the stellar population. In section 4 we model the star formation in the disk with a Schmidt law and a gas infall law, assuming the same model of the chemical evolution of the Galaxy as \\scite{Matteucci+89}. Finally, in section 5 we compute synthetic spectra and photometric properties of the stellar population in the disk, using the stellar evolution tracks of JMSTAR15, stellar atmosphere models by \\scite{Kurucz_92} and J{\\o}rgensen (private communication) and the chemical evolution models built for LSBs (see section 4). We discuss the results in section 6. The most important results of this work are: \\begin{enumerate} \\item Observed colour profiles, chemical abundances and surface brightness profiles for LSBs are well fitted if they are assumed to have a spin parameter for its halo higher than HSBs. \\item LSBs are not young objects, as often claimed in the literature, since their colours are well fitted by old ($> 7$ Gyr) stellar populations. \\item There is discrepancy between the photometric age of the galaxies and the age of formation of their halo, which indicates that the star formation can start about 2~Gyr before the halo is formed. This discrepancy is reduced to 1~Gyr if the Hubble constant is assumed to be $H_0=65$~kms$^{-1}$Mpc$^{-1}$ and completely removed if the Universe is open ($\\Omega < 0.3$) or has a significant vacuum energy contribution ($\\Lambda > 0.6$). \\item The earliest stellar populations of LSBs could be present in the high redshift universe in a similar proportion to HSBs as they are now. Their colours at high redshift (as the colours of all disks) are about 1~mag bluer than at low redshift. \\end{enumerate} In a previous paper (\\pcite{Padoan+Jimenez+Antonuccio97a}) we have shown that LSBs are not necessarily young and un-evolved systems (e.g. \\pcite{McGaugh+Bothun94,deBlok_phot+95}). We used the simplest possible model, a burst of star formation, to determine a lower limit to the age of LSBs studied by \\scite{deBlok_phot+95}. In the present work we considerably improve our previous model because we use a more realistic continuous star formation process, we include self--consistently the chemical evolution, and we connect the disk model to the cosmological scenario using the spherical collapse model (\\pcite{Gunn_Gott_72}). This more detailed model confirms our previous result that the blue LSB disks in the sample by \\scite{deBlok_phot+95} are not un-evolved objects collapsed at late times from low initial over-densities (\\pcite{McGaugh+Bothun94,Mo+94}), but rather normally evolved disk galaxies. Throughout the paper the present day value of the Hubble constant is assumed to be $H_0=75$~kms$^{-1}$Mpc$^{-1}$. ", "conclusions": "\\subsection{The Nature of LSBs} \\begin{figure*} \\centering \\leavevmode \\epsfxsize=1.4 \\columnwidth \\epsfbox{lsbabund.eps} \\caption[]{The redshift evolution for some elements is plotted. The top-left panel shows the evolution of $[O/H]$ and is compared with \\scite{McGaugh_oxigen94} measures. Our model reproduces perfectly the peak in $[O/H]$ among LSBs found by \\scite{McGaugh_oxigen94} at $[O/H]=-0.9$. The top-right panel shows the analogous to the previous one but for $[Zn/H]$ compared with \\scite{Pettini+97} measures. It transpires from the figure that the predicted $[Zn/H]$ for LSBs is consistent with the spread found for DLAs.} \\label{f17} \\end{figure*} In this work we have shown that the photometric properties of the bluest galaxies in the sample by \\scite{deBlok_phot+95} can be reproduced assuming that they form and evolve as normal disk galaxies, with relatively high spin parameter. That the spin parameter alone could explain the low surface brightness was already shown by \\scite{Dalcanton_disc+97}, but they assumed a given $M/L$ ratio. Here we have strengthened the point by also explaining the colours, the colour gradients, and the chemical abundances. We have also provided synthetic spectra that could be compared with the observations, in the effort to prove that LSB disk galaxies are indeed normal galaxies with high spin parameter. The stellar populations of LSBs are rather evolved, especially in the central part of their disks. Even the bluest galaxies in the LSB sample by \\scite{deBlok_phot+95} are older than 7~Gyr, and typically star formation in these galaxies starts about 9~Gyr ago. This confirms the results of \\scite{Padoan+Jimenez+Antonuccio97a}, where the galaxies in the same sample were found to be at least 7~Gyr old. LSB and HSB disks are therefore hosted in similar dark matter halos that differ only for their spin parameter, and they evolve in a similar way, apart from the fact that LSB disks are less concentrated than HSB disks. This means that the present day (or low redshift) abundance of LSBs relative to HSBs should be about the same at any redshift. Moreover, the conclusions about the epoch of star formation and halo formation (see the next section), and about the colour evolution must be valid for disk galaxies in general. In particular, Fig.~\\ref{f11} shows that the colour evolution of LSBs is very strong. The U-B, B-V and V-I colours of the central part of the disk at high redshift are about 1~mag bluer than they are at low redshift, and the B-R colour even 1.5~mag bluer. This must be a property of disks in general, and not only of LSB disks. \\subsection{Star Formation and Galaxy Formation} We have shown in section 2.1 that the size of LSB disk galaxies is such that their halos are formed in the redshift range $0.1 7$ Gyr) stellar populations. \\item There is discrepancy between the photometric age of the galaxies and the age of formation of their halo, which indicates that star formation can start about 2~Gyr before the halo is formed. This is perfectly acceptable in hierarchical models of galaxy formation, and, as discussed in the text, the discrepancy can be reduced or removed if the cosmological model we have assumed is incorrect. If an open ($\\Lambda$) cosmology had been adopted, the above discrepancy would have been completely removed. \\end{enumerate}" }, "9804/astro-ph9804339_arXiv.txt": { "abstract": "The variable star population of the galactic globular cluster NGC 1851(C0512-400) has been studied by CCD photometry, from observations made in the B, V and I bands during 1993-1994. Light curves are presented for 29 variables, seven of which are new discoveries. The behavior of the RR Lyraes in the period-temperature diagram appears normal when compared to clusters which bracket the NGC 1851 metallicity. Reddening and metallicity are re-evaluated, with no compelling evidence being found to change from the values of $E(B-V) = 0.02$ and $[Fe/H] = -1.29$ (Zinn scale) adopted in recent studies of the cluster. Photometry is provided for stars in an annulus with radii 80 and 260 arcsec centered on NGC 1851. To at least $V=18.5$ there is excellent agreement with the extensive earlier photometry for the brighter NGC 1851 stars, with systematics less than 0.02 mag in all colors. Instability strip boundary positions for several clusters shows a trend for the red boundary to move to redder colors as the metallicity increases. keywords: globular clusters: individual (NGC1851) - RR Lyrae variable ", "introduction": "The galactic globular cluster (GC) NGC 1851 (C0512-400) is rich, centrally-condensed and belongs to the small group of clusters which display bimodal horizontal-branch (HB) morphology, defined (Catelan et al 1998) as having fewer RR Lyrae stars than either blue or red HB stars. Canonical theory, that is considering a GC as a population characterized by a single age, constant abundance and a red-giant branch (RGB) mass loss parameter that has a narrow Gaussian distribution (typically $\\sigma_M \\sim 0.02 M_{\\sun}$), cannot explain such unusual HB morphology. In order to account for the bimodality, attention has focussed recently on scenarios which can alter the mass loss parameter, such as tidal stripping of red-giant envelopes in dense environments, rapid rotation, stellar encounters, and binary interactions. Sosin et al (1997) discuss these various options in the context of the most extreme example known of a GC with bimodal HB, NGC 2808, which displays a blue HB with multiple gaps that extends to below the main sequence turnoff in the $V, B-V$ color-magnitude diagram (CMD), $M_V \\sim 5$. They conclude that none of the present explanations are a satisfactory match to the observations. However, Sweigart \\& Catelan (1997) have modeled the unusual HB morphology of the metal rich GC's NGC 6338 ($[Fe/H] = -0.60$) and NGC 6441 ($[Fe/H = -0.53$) for which Rich et al (1997) have obtained CMD's using the Hubble Space Telescope. Both clusters have HB's which slope upwards (brighter) with decreasing $B-V$, and have extended blue tails. Models with high helium abundance, rapid rotation, and helium mixing into the envelope are all able to produce a sloping HB morphology, and sometimes a bimodal distribution. The helium mixing alternative is particularly interesting given the observed heavy-element abundance variations in globular cluster red-giant stars (Kraft 1994). Mixing deep enough to produce enhanced aluminium, as observed in some stars, will also dredge up helium. Extensive deep mixing might be expected to destroy the sharp boundary corresponding to the deepest penetration of the convective zone, and thus prevent the observational pile-up of stars on the RGB near the level of the HB. NGC 1851 in fact appears to have quite a prominent such clump, thus suggesting that deep mixing has not taken place on the RGB at a luminosity less than that of the clump. Notwithstanding, the number of possible options still available to explain the peculiar NGC 1851 HB is considerable. The CMD of NGC 1851 has most recently been studied by Walker (1992) (hereafter W92) in the $B$ and $V$ bands, in the UV by Parise et al (1994) and in the $V$ and $I$ bands by Saviane et al. (1997) (hereafter S97), where references to earlier work can be found. In both optical studies the bimodal HB is interpreted as a consequence of differing efficiencies of mass loss as the stars evolve up the RGB. W92 suggested that a unimodal mass distribution might be able to produce a bimodal HB stellar distribution when the detailed shape of the evolutionary tracks was taken into account, however S97 do not find good agreement when comparing with the Bertelli et al. (1994) tracks. They prefer a bimodal mass loss distribution, and indeed find some evidence to suggest that the radial distributions of the blue and red HB stars differ, pointing towards some as yet unexplained interaction between the dynamical evolution and the stellar evolution of these stars. On the other hand, Catelan et al (1998), using updated (Sweigart 1997) Sweigart and Gross (1976, 1978) models, find they can reproduce the NGC 1851 HB morphology with a unimodal, albeit very wide, mass distribution having characteristics $ = 0.665 M_{\\sun}$ and $\\sigma_M = 0.055 M_{\\sun}$. Stetson et al. (1996) and Sarajedini et al. (1997) review the question of the relative ages of globular clusters, to reach very different conclusions. In both cases the HB level of NGC 1851 (W92) is used to link the second-parameter pair NGC 288 and NGC 362. Critical to these arguments is the V magnitudes of the reddest BHB stars and the RR Lyraes; the latter will be provided here for the first time. Lying in the region of the HB which is sparsely populated, the RR Lyraes may provide important clues to help explain the reason for the bimodal HB, given the constraints that the pulsation properties place on stellar evolutionary status. There have been suggestions (Catelan 1997) that the 1851 variables are peculiar with respect to their behavior in the period-temperature diagram and that the photographic studies, as detailed below, also show that several of the RRab stars have light curve amplitudes near 2 magnitudes, much larger than normal. Sawyer Hogg (1973) lists 10 variables in NGC 1851, from discoveries by Bailey (1924) and Laborde and Fourcade (1966). Preliminary periods for some of these stars, and an additional four new discoveries, resulted from a short observing campaign by Liller (1975) using photographic photometry at the CTIO 1.0-m and 1.5-m telescopes. She noted that V2 and V8 appeared to be constant, and V9 was very red. Wehlau et al. (1978) (hereafter W78) measured an additional 57 plates, almost all taken with the 1.0-m Swope telescope at Las Campanas, and analysed them along with the Liller (1975) plates. Periods were derived and light curves presented for a total of 19 RR Lyrae stars, the mean period of the RRab stars being 0.573 days, and the ratio of the number of RRc to RRab variables was found to be 0.36, both are values typical of an Oosterhoff type I system. The photometric zeropoint calibration for these observations was very uncertain, due to the lack of a definitive photometric sequence in the field. Stetson (1981) found four additional variables and two apparently constant stars lying in the instability strip. Wehlau et al. (1982) (hereafter W82) studied these stars using their original plate material supplemented with another 18 plates taken in 1970, whereupon three stars were found to be RR Lyraes, a fourth was classified as a probable field W UMa star, while a fifth is a red variable. W78 and W82 determined accurate periods and approximate mean $$ magnitudes for all the stars they identified as RR Lyraes, classifying 15 stars as RRab and seven as RRc. The light curves display the $0.1-0.2$ mag scatter typical for the photographic technique. No modern studies of the NGC 1851 variables appear in the literature, however S97 identify an additional seven candidate RR Lyrae variables from a comparison of their photometry with that of W92, selecting stars with deviant photometry and $V \\sim 16, V-I \\sim 0.4$. Here we present new CCD photometry for the NGC 1851 RR Lyrae variables and compare with results for RR Lyraes in other GC's in this program (IC 4499, M68, M72, NGC 6362). ", "conclusions": "" }, "9804/astro-ph9804027_arXiv.txt": { "abstract": "The Hubble Deep Field-South was chosen to have a QSO (RA 22:33:37.6 Dec $-$60:33:29 J2000 and B=17.5) in the field to allow for studies of absorption systems intersecting the sight line to the QSO. To assist in the planning of HDF-S observations we present here a ground-based spectrum of the QSO. We measure a redshift of $z=2.24$ for the quasar and find associated absorption in the spectrum at $z=2.204$ as well as additional absorption features. ", "introduction": "Unlike the original Hubble Deep Field, the Hubble Deep Field South (HDF-S) was chosen specifically to contain a $z>2$ QSO suitable for studying the relationship between the high redshift galaxies identified in the HDF-S and the absorption lines in the spectrum of the HDF-S QSO. The QSO was found on a UK Schmidt Telescope objective prism plate scanned by Mike Irwin using the Automated Plate Measuring facility in Cambridge analysed by Paul Hewett and then confirmed by observations at the Anglo-Australian Telescope (Boyle 1997). To aid future observations of the HDF-S we present here a low-resolution spectrum of the QSO; the data are available at http://bat.phys.unsw.edu.au/$\\sim$kms/hdfs/. ", "conclusions": "" }, "9804/astro-ph9804211_arXiv.txt": { "abstract": "We use the measures of Li and rotational velocities in F Hyades stars to assess the role of the wind-driven meridian circulation and of shear turbulence in the transport of angular momentum in stars of different masses. Our models include both element segregation and rotation-induced mixing, and we treat simultaneously the transport of matter and angular momentum as described by Zahn (1992) and Maeder (1995). We show that the hot side of the Li dip in the Hyades is well explained within this framework, which was also successfully used to reproduce the C and N anomalies in B type stars (Talon et al. 1997). On the cool side of the dip, another mechanism must participate in the transport of angular momentum; its efficiency is linked to the depth of the surface convection zone. That mechanism should also be responsible for the Sun's flat rotation profile. ", "introduction": "During the last decade, special efforts have been devoted to improve the description of the mixing processes related to stellar rotation. The most recent works (see for example Pinsonneault et al. 1989, Zahn 1992, Maeder 1995, Talon \\& Zahn 1997) describe the evolution of the internal distribution of angular momentum in a self-consistent manner under the action of meridional circulation and of shear turbulence. The mixing of chemicals is then linked directly to the rotation profile, whereas previous studies made use merely of a parametric relation between the turbulent diffusivity and the rotational velocity (cf. e.g. Schatzman et al. 1981, Zahn 1983). Such a self-consistent treatment was applied successfully by Talon et al. (1997) in the study of a 9 M$_{\\odot}$ star, modeling the transport of angular momentum by the meridional circulation as a truly advective process. The only assumption in this theory is that the turbulence sustained by the shear is highly anisotropic and relies on two free parameters; the first one describes the magnitude of the horizontal shears (cf. Zahn 1992) and the second one, the erosion of the restoring force due to both the thermal and the mean molecular weight stratifications (cf. Maeder 1995, Talon \\& Zahn 1997). These authors reproduce the slight abundance anomalies measured in B stars by Gies \\& Lambert (1992). They also show that the widening of the main sequence, which is generally attributed to convective overshooting in massive stars, may be due to the rotational mixing present in stars having a ``typical'' velocity for the spectral type considered. Concerning low-mass stars, it has been shown that the hydrodynamical models relying on meridional circulation and shear fail to reproduce the solar rotation profile given by the helioseismic observations (Brown et al. 1989, Kosovichev et al. 1997): at the solar age, those models still have large $\\Omega$ gradients which are not present in the Sun (see Chaboyer et al. 1995 and Matias \\& Zahn 1997). That conclusion has been reached independently by two different groups, using different descriptions for the transport processes. On one hand, the Yale group computed the evolution of angular momentum in low mass stars with a simplified description of the action of the meridional circulation which was considered as a diffusive process rather than as an advective process. The whole evolution of momemtum and chemicals was then due to diffusion only, with a free parameter that had to be calibrated to differentiate the transport of the passive quantities with respect to that of vectorial ones. Pinsonneault et al. (1990) were then able to reproduce the surface Li abundances for low-mass cluster stars (with effective temperature lower than 6500K). However, they obtained large rotation gradients within these stars which are excluded by helioseismology (Chaboyer et al. 1995). On the other hand, Matias \\& Zahn (1997) performed a complete study for the evolution of the Sun's angular momentum, where they took into account the advective nature of the meridional circulation. They also concluded that meridional circulation and shear turbulence are not efficient enough to enforce the flat rotation profile measured by helioseismology. These results indicate that another process participates in the transport of angular momentum in solar-type stars, while the so-called wind-driven meridional circulation (Zahn 1992) is successful in more massive stars. In order to study the transition between solar-type and more massive stars and to identify the mass range for which the present description for the transport of angular momentum and chemicals relying only on rotation fails, we propose to use the measures of lithium and rotational velocities in galactic cluster stars. We first review the observations of lithium abundances and rotation in the Hyades main-sequence stars, and summarize the difficulties of the various models proposed so far to explain the Li dip in F stars (\\S 2). We recall the equations that describe the evolution of angular momentum due to meridian circulation and shear turbulence as well as the associated transport of chemicals (\\S 3). We study the impact of rotational mixing on the lithium abundance in galactic cluster F stars, and compare this to the observations. Our models include both element segregation and rotation-induced mixing, and we treat simultaneously the transport of matter and angular momentum. The internal rotation profile thus evolves completely self-consistently under the action of meridional circulation as described by Zahn (1992) (see also Matias et al. 1997), and of shear stresses which take into account the weakening effect of the thermal diffusivity, as was first shown by Townsend (1958) (\\S 4). We show that the blue side of the lithium dip is well reproduced within this framework, and that the process responsible for the shape of the solar rotation profile should become efficient only for stars on the cool side of the Li dip, where the external convection zone is thick enough. By achieving efficiently momentum transport, the global effect of this process would be to reduce the mixing due to the rotational instabilities in stars with effective temperature lower than $\\sim$ 6500K. The most likely candidates for this transport process are the gravity waves generated by the external convection zone (Schatzman 1993; Zahn et al. 1997; Kumar \\& Quataert 1997) and the large-scale magnetic field which could be present in the radiative interior (Charbonneau \\& MacGregor 1993; Barnes et al. 1997). ", "conclusions": "Assuming rapid rotation and using a self-consistent description for the transport of angular momentum and of chemicals by meridional circulation and shear instabilities (cf. Zahn 1992, Talon \\& Zahn 1997), Talon et al. (1997) successfully explained the C and N anomalies observed in some B stars. At the same time, it was shown (Matias \\& Zahn 1997) that this description applied to the transport of angular momentum in the Sun is incomplete, leading to large $\\Omega$ gradients which are not observed. Another transport mechanism must thus be invoked in low mass stars. At this point, 2 questions remain : firstly, the nature of that transport mechanism has to be determined unambiguously and secondly, the location of the transition between the regime which is relevant for massive stars and the one which is relevant for low mass stars has to be identified. In this paper, we addressed that second question. We presented numerical calculations of Li destruction due to rotational mixing using the {\\it same} description as Talon et al. used for more massive stars and the same free parameters. We showed that this clearly reproduces the hot side of the Li dip. Let us recall that the destruction of lithium is then due solely to rotational mixing enhanced by the spin down of the outer layers. Stars hotter than 7000 K also undergo rotational mixing, but it is much milder due to the weak differential rotation. The rise of Li abundances on the right side of the dip is not explained within this framework. We propose that it is linked to the appearance of another transport mechanism for angular momentum which reduces the magnitude of the meridional circulation and shears, leading to the observed diminution of Li destruction on the red side of the Li dip. This mechanism is known to occur in the Sun where it is responsible for the flat rotation profile." }, "9804/astro-ph9804161_arXiv.txt": { "abstract": "The Berkeley spectrograph aboard the ORFEUS telescope made its second flight on the 14-day ORFEUS-SPAS II mission of the Space Shuttle {\\it Columbia} in November/December 1996. Approximately half of the available observing time was dedicated to the Berkeley spectrograph, which was used by both Principal and Guest Investigators. The spectrograph's full bandpass is 390--1218 \\AA; here we discuss its in-flight performance at far-ultraviolet (FUV) wavelengths, where most of the observations were performed. The instrument's effective area peaks at 8.9 $\\pm$ 0.5 cm$^2$ near 1020 \\AA, and the mean spectral resolution is 95 km s$^{-1}$ FWHM for point sources. Over most of the spectral range, the typical night-time background event rate in each spectral resolution element was about 0.003 s$^{-1}$. Simultaneous background observations of an adjacent blank field were provided through a secondary, off-axis aperture. The Berkeley spectrograph's unique combination of sensitivity and resolution provided valuable observations of approximately 105 distinct astronomical targets, ranging in distance from the earth's own moon to some of the brightest AGN. ", "introduction": "The German spacecraft Astro-SPAS, a deployable platform designed to meet the technical performance demands of astronomical payloads and similar scientific instruments, comprised the primary payload aboard shuttle mission STS-80 ({\\it Columbia}). On this, its third flight, the platform carried a trio of far-ultraviolet instruments: two independent spectrographs within the 1 meter diameter ORFEUS telescope (Grewing et al. 1991) and the IMAPS objective-grating spectrograph (Jenkins et al. 1996). All three had flown on the Astro-SPAS' 5-day maiden voyage in September of 1993, but improvements in instrument performance, and the critical need for additional observation time, motivated a reflight. A photograph of the payload is shown in Plate~1. With few exceptions, targets suitable for ORFEUS were too faint for IMAPS, so no attempt was made to coalign these instruments closely. Within the ORFEUS telescope a flip mirror was employed to direct the optical beam to one spectrograph or the other. Hence in general only one of the three instruments was operated at a time. The available observing time was shared equally between Guest Investigators selected by peer review and the Principal Investigator teams who had provided the instruments. Flight operations were directed from a control complex at the Kennedy Space Center. The general design of the Berkeley spectrograph has been discussed previously (Hurwitz \\& Bowyer 1986, 1996). We changed the instrument between missions only by overcoating of two of the four diffraction gratings (including the far ultraviolet grating) with silicon carbide, introducing the multiple apertures discussed below, and modifying the detector electronics to improve the imaging at high count rates. We did not recoat the KBr photocathode on the microchannel plate detectors; the delay-line anode detector systems are discussed in Stock et al. (1993). In this work we report on the performance and calibration of the spectrograph during the ORFEUS-SPAS II mission and the instrumental effects of interest to Guest Investigators and other users of the extracted data products. ", "conclusions": "The Berkeley spectrograph aboard the ORFEUS telescope offers a unique and important combination of spectral resolution and effective area in the comparatively unexplored far-ultraviolet wavelength band. During the ORFEUS-SPAS II mission in November/December 1996, Principal and Guest Investigators utilized the spectrograph to observe some 105 astronomical targets. These data will enter the public domain in early to mid 1998. Lyman/FUSE is not far from launch, and will offer a much higher spectral resolution and sensitivity for point sources. However, near-trivial modifications would enable the Berkeley spectrograph to achieve a significantly superior performance for studies of extended emission. New replicas of the current gratings and SiC overcoating of the primary mirror would allow the spectrograph to achieve 2 \\AA\\ slit-limited resolution through a 45 $\\times$ 420\\arcsec\\ slit and an effective area of $\\sim$ 50 cm$^2$ across the 900 -- 1200 \\AA\\ band. The angular resolution along the slit length would be better than 1 \\arcmin. Such an instrument would be highly desirable for studies of intracluster gas, galaxies and their halos, supernova remnants, and other extended objects. At time of writing, there are no specific plans for a third flight of the payload." }, "9804/astro-ph9804283_arXiv.txt": { "abstract": "In this paper we propose a new statistic capable of detecting non-Gaussianity in the CMB. The statistic is defined in Fourier space, and therefore naturally separates angular scales. It consists of taking another Fourier transform, in angle, over the Fourier modes within a given ring of scales. Like other Fourier space statistics, our statistic outdoes more conventional methods when faced with combinations of Gaussian processes (be they noise or signal) and a non-Gaussian signal which dominates only on some scales. However, unlike previous efforts along these lines, our statistic is successful in recognizing multiple non-Gaussian patterns in a single field. We discuss various applications, in which the Gaussian component may be noise or primordial signal, and the non-Gaussian component may be a cosmic string map, or some geometrical construction mimicking, say, small scale dust maps. ", "introduction": "Current theories of structure formation may be roughly divided into two classes: active and passive perturbations. According to the inflationary paradigm, quantum fluctuations in the very early universe are produced during a period of inflation \\cite{cosmcross} and grow to become classical density perturbations \\cite{lidlyth,bardeen}. These perturbations evolve linearly until late times when the overdensities become galaxies. Perturbations due to inflation are called passive because they are seeded at some initial time near the Planck time and then evolve `deterministically', or linearly. They leave their imprint on the CMB at last scattering \\cite{pyu,nature,HS1,HS2,richard,1.3}. Although this is not strictly necessary, in most cases the fluctuations in the CMB temperature due to inflationary perturbations form a Gaussian random field \\cite{bondefst,bardstat}. There is another class of theories of structure formation, topological defects caused by phase transitions in the early universe \\cite{Kib,vv}. Perturbations caused by defects are known as active perturbations \\cite{mafc,coher} since they are continually being seeded by an evolving network of defects through the history of the universe. The fluctuations in the CMB in defect models have been found to be non-Gaussian \\cite{phases}, even though the extent and strength of this non-Gaussianity is still far from clear. Recently a wide class of defect models have been shown to be in conflict with current data \\cite{neil,abr,abr2} but some viable active models still remain \\cite{mimic,durrer,abrw}. In fact, some of the most interesting \\cite{abr3,shellard} of these (based on cosmic strings) require a non-zero cosmological constant of the sort that currently favoured by supernova experiments \\cite{sn1}. Thus it is important to be able to distinguish Gaussian from \\ngn fluctuations. Many different tests are being tried \\cite{Kogut,XLuo,fermag,cumul,gorski,sergei}, adapted to different experimental settings, and types of signal. Our statistic is designed to be used for small fields where one or more distinctive shapes are obscured by an extra Gaussian component, and so are not visible in real space. It is our experience in previous work \\cite{fermag} that in such situations standard statistics, based in real space, fail to recognize the non-Gaussianity of the signal. The idea is to study the statistical properties of map derivatives which are only sensitive to a given scale. In this way we may separate out different scales, some of which may have Gaussian or nearly Gaussian fluctuations, some very non-Gaussian. Although this scale filtering may be achieved using the wavelet transform \\cite{cumul,pando}, in this paper we choose to use the Fourier transform as our scale filter. One problem with this approach is that the Fourier transform is a global transformation, and therefore can only recognize non-Gaussian structures globally. It may therefore offer a rather contorted description for complicated networks made up of essentially simple objects. We will however perform a transformation over the Fourier modes: a second Fourier transform, in angle, for modes in each ring in Fourier space. We will argue that by doing so we are able to recognize mostly features of individual objects, and so bypass this problem. This operation returns a set of quantities which are blind to the random orientations of individual objects. Unfortunately their random positions still affect our statistic, so there will be a limit (albeit less stringent) on the number of objects in the field before our statistic becomes confused. The plan of this paper is as follows. In Section II we look at some of the ideas involved in thinking about non-Gaussianity. In Section III we define our statistic and look at some of its properties and motivations. In Section IV we apply our statistic to practical situations in which subtle non-Gaussian signals are present. We consider non-Gaussian signals corrupted by the presence of a Gaussian process, which can be noise or primordial signal, and which dominates on all but a narrow band of scales. We consider two types of non-Gaussian signal. We consider CMB maps obtained from string simulations, implementing the algorithms in \\cite{av}. We also consider geometrical constructions mimicking, say, small scale dust maps. We show how such maps look very Gaussian, but fail to confuse our statistic. ", "conclusions": "We have introduced a statistic to look for \\ng in the cosmic microwave background. It is designed to pick out \\ngn features which are superimposed on Gaussian fluctuations and which are therefore not visible in real space. Since some scales may be more \\ngn than others we choose to separate them out. It is our experience in previous work \\cite{fermag} that in such situations standard statistics, based in real space, fail to recognize the non-Gaussianity of the signal. Our statistic is naturally tailored for interferometric experiments, which make measurements in Fourier space. Indeed the algorithm exposed above could easily be turned into a data analysis package operating over visibilities. One of the problems with Fourier space statistics is that they recognize only global shapes. Therefore they become very ineffective when many individual structures are present. By taking another Fourier transform, in angle, over modes in a given ring in Fourier space, we can factor out orientations, and be only sensitive to an average shape of individual structures. Their random positions, on the other hand, will eventually make our statistic very ineffective as the number of objects becomes large. We found that although we have improved on previous work, still there is a limit on how many individual structures there may be in the field before the statistic fails to pick out their non-Gaussianity. Finally, some comments are in order regarding the applications we have considered. The Gaussian component we have considered may include noise, if it has the right power spectrum. Hence we have already considered the effects of a simplified form of noise: it will merely reduce the band of scales where the non-Gaussian signal dominates, typically providing it with an upper boundary. An exception is the case of non uniform noise in the u-v plane, present in most interferometer experiments. Non uniform noise in the Fourier domain will look non-Gaussian, for it is a Gaussian process which is not isotropic or translationally invariant. Therefore an extra element of confusion, not dealt with in this paper, will appear. Regarding the non-Gaussian component we have considered two types of signal. We have looked at some of the properties of the statistic when it is applied to a field containing one or more distinct shapes. The geometrical constructions presented are somewhat reminiscent of the jagged structures present in small scale dust maps. Experimenting on fields containing many structures we saw our statistic rapidly become blind to their non-Gaussianity. However interferometer fields are often small. They would therefore contain only a few, but more than one, of these structures. This is something our statistic can cope with, unlike previous Fourier space based statistics. We have also studied our statistic when applied to cosmic string maps. We found that some individual small fields showed \\ng in the signal. If the results were averaged over many fields the effect of \\ng on $|F(\\beta)|$ would not show. Once again this strategy is ideal for interferometers, for which combining small fields into a large field is in fact a rather awkward operation." }, "9804/astro-ph9804230_arXiv.txt": { "abstract": " ", "introduction": "La d\\'ecouverte en 1965 par Penzias et Wilson du fond de rayonnement cosmologique diffus \\`a 3 K a d\\'efinitivement fait du mod\\`ele du Big Bang chaud le cadre g\\'en\\'eral de la cosmologie moderne. Ces trois derni\\`eres d\\'ecennies ont \\'et\\'e marqu\\'ees par de nombreux progr\\`es tant th\\'eoriques qu'observationnels qui n'ont fait que confirmer sa validit\\'e, et de pr\\'eciser petit \\`a petit les d\\'etails de ce sc\\'enario. Parmi les avanc\\'ees notables de ces derni\\`eres ann\\'ees, citons pour m\\'emoire: \\begin{itemize} \\item La r\\'ealisation de grands catalogues de galaxies. On a maintenant des catalogues tri-dimensionnels qui contiennent des milliers d'objets. Ces catalogues permettent de faire une v\\'eritable cosmographie de l'univers local. De nouveaux moyens sont apparus r\\'ecemment pour compl\\'eter ce type de catalogues, la d\\'etermination des champs de vitesse cosmiques, permet ainsi d'acc\\'eder \\`a des informations cin\\'ematiques tr\\`es pr\\'ecieuses. Et depuis la fin des ann\\'ees 80, un nouveau moyen d'investigation est en train d'\\'emerger, il s'agit des cartes de distorsion gravitationnelle. Elles permettent de visualiser les lignes de potentiel de la masse projet\\'ee. \\item le mod\\`ele de Mati\\`ere Noire Froide (CDM pour Cold Dark Matter) qui est apparu au d\\'ebut des ann\\'ees 80 et qui a servi de point de r\\'ef\\'erence (\\`a d\\'efaut de devenir un mod\\`ele standard) pour tous les travaux sur le probl\\`eme de la formation des grandes structures. \\item Le d\\'eveloppement des th\\'eories inflationnaires. C'est encore un terrain tr\\`es sp\\'eculatif, mais c'est le lieu tr\\`es excitant o\\`u les concepts de la physique des hautes \\'energies rencontrent des exigences observationnelles de plus en plus fiables. \\item Enfin, la d\\'etection en 1992 des fluctuations de temp\\'erature du fond de rayonnement cosmologique par l'exp\\'erience satellitaire COBE/DMR (Smoot et al. 1992) a marqu\\'e un tournant pour la cosmologie: pour la premi\\`ere fois on avait une preuve directe de l'origine des grandes structures de l'univers. \\end{itemize} Dans ce cours je vais principalement m'int\\'eresser au probl\\`eme de la formation des grandes structures. Quelques ouvrages de r\\'ef\\'erence, \\begin{itemize} \\item Relativit\\'e G\\'en\\'erale: Weinberg, {\\it Gravitation and Cosmology}, 1972; Landau et Lifschitz, {\\it Classical Theory of Fields}, 1975 \\item Univers primordial: Kolb et Turner {\\it The Early Universe}, 1990, mais il est peu pr\\'ecis; revue de Brandanberger, {\\it Inflation and Cosmic Strings: two Mechanisms for producing Structure in the Universe}, Int. J. Mod. Phys. A2 : 77, 1987. \\item Inflation: Linde, {\\it Particle Physics and Inflationary Cosmology}, 1990; Liddle et Lyth, {\\it The Cold Dark Matter density perturbation}, 1993, Physics Reports, 231, 1; Lidsey et al. {\\it Reconstructing the inflaton potential - an overview}, 1997, Reviews of Modern Physics, 69, 2 \\item Formation des grandes structures: Peebles {\\it The Large Scale Structure of the Universe}, 1980; {\\it Principle of Physical Cosmology}, 1993; V. Sahni \\& P. Coles, {\\it Approximation Methods for Non-linear Gravitational Clustering}, 1995, Physics Reports, 262, 1 \\end{itemize} ", "conclusions": "Dans ce petit tour d'horizon de la cosmologie j'ai \\'et\\'e loin d'\\^etre exhaustif. Dans le domaine de l'\\'evolution des grandes structures je n'ai pas mentionn\\'e bon nombre de domaines de recherche en plein d\\'eveloppement, comme \\begin{itemize} \\item L'\\'etude des flots cosmiques \\`a grande \\'echelle, et leur utilisation pour mesurer les param\\`etres cosmologiques; \\item L'\\'etude des amas de galaxies et de leur contenu aussi bien en mati\\`ere noire (\\'etudes dynamique, reconstruction de masse par effets de lentilles gravitationnelles..), qu'en mati\\`ere baryonique (galaxies, rayonnement X..); \\item L'\\'etude des absorbants et des objets de grand $z$; \\item ... \\end{itemize} Du point de vue du physicien th\\'eoricien un domaine de recherche de pr\\'edilection est inconstablement la cosmologie primordiale, lieu de rencontre entre entre la physique des hautes \\'energies et la cosmologie observationnelle. Cependant la formation des grandes structures recelle un certain nombre de probl\\`emes ouverts digne d'int\\'er\\^et, \\begin{itemize} \\item La th\\'eorie des perturbations est loin d'avoir livr\\'e tous ses secrets. En particulier il serait extremement utile de comprendre les corrections en boucles. \\item De nombreux aspects du r\\'egime non-lin\\'eaire ne sont pas compris. Par exemple on ne sait pas d\\'ecrire la transition vers le r\\'egime multiflots qui conduit \\`a la virialisation de la mati\\`ere dans les puits de potentiel. Il serait aussi tr\\`es int\\'eressant de pouvoir exhiber une solution explicite des \\'equations dynamiques dans le r\\'egime nonlin\\'eaire (m\\^eme si ce n'est qu'une forme assymptotique). \\item L'exploitation des catalogues tridimensionnel ou bidimensionnel notamment avec la mise en \\'evidence de propri\\'et\\'es non-Gaussiennes n'est pas encore optimale. La question se pose par exemple pour les cartes de distorsion gravitationnelle. \\item La relation entre baryons et mati\\`ere noire est un domaine qui n'a pratiquement pas \\'et\\'e explor\\'e. On ne dispose pour l'instant que d'exp\\'eriences num\\'eriques pour essayer de comprendre ce qui se passe. \\end{itemize}" }, "9804/astro-ph9804006_arXiv.txt": { "abstract": "The Lyman edge at 912 {\\AA} is an important diagnostic region for studying quasi-stellar objects (QSOs). In particular, it reveals a great deal about the physical conditions within the atmospheres of accretion disks, a ubiquitous component of QSO theories. A robust prediction of accretion disk models is a significant polarization due to electron scattering just longward (in wavelength) of the Lyman edge because of the wavelength dependence of the Hydrogen absorption opacity. Observations of the Lyman edge regions of QSOs have shown scant evidence for the predicted features---few QSOs show the broad, partial Lyman edges expected to be common according to most theories, and none show the high polarizations expected longward of the Lyman edge. Still, polarization spectra of a small number of QSOs have shown a rising polarization (up to 20\\%) at wavelengths {\\it shortward} of the Lyman edge. We have now doubled our sample of intermediate-redshift QSOs observed with the {\\it HST/FOS} spectropolarimeter to determine the amount of polarization on both sides of the Lyman limit. For this new sample of six objects, polarizations are low and mostly consistent with zero below the Lyman edge. Another important result of the new data is that it strengthens the conclusion that quasars are generally not polarized significantly just longward of the Lyman edge at $\\sim$1000\\AA\\/. There is no significant statistical wavelength dependence to the polarization longward of the Lyman edge indicating that simple plane-parallel atmospheres with scattering-dominated opacity are not significant sources of UV flux in quasars. ", "introduction": "One of the fundamental components of most theories of quasi-stellar objects (QSOs) is an accretion disk. As gas is fed into the central regions of the QSO, residual angular momentum causes the gas to naturally settle into a disk. While most theories predict the formation of such a disk, few address the observational consequences of the disk models in detail. These theoretical studies have found that the Lyman edge at 912 {\\AA} is a powerful diagnostic feature for the physical characteristics of the disk. The simplest disk models (quasi-static with viscous dissipation at large optical depth) predict Lyman edges in either emission or absorption, depending on the viewing angle and the physical details of the disk atmosphere. Such Lyman edges would be broadened by rotation of the disk and by general relativistic effects as light passes close to the central black hole. In most QSOs such edges are not seen, although Koratkar, Kinney, \\& Bohlin (1992) found a small number of candidate ``partial edges'' in IUE data. A second disk signature is the linear polarization, $P$, of the continuum from the disk. A purely scattering atmosphere will produce high polarization perpendicular to the disk axis. Again, this signature is not seen in any QSOs; in fact generally QSOs show {\\it low} optical polarization {\\it parallel} to the inferred disk axis. Laor, Netzer, \\& Piran (1990) attempted to show that a disk atmosphere should have significant absorptive opacity, and thus can produce dramatically lower optical polarization (albeit still perpendicular to the disk axis and thus inconsistent with the observations). A more robust prediction according to their work, however, is a rise in polarization with decreasing wavelength from the optical into the UV. Just longward of the Lyman edge $P$ is highest, since it is at these wavelengths that scattering best competes with absorption. Just shortward of the Lyman edge, as the absorption opacity increases, $P$ should drop again. According to Laor \\etal, this polarization signature should appear even when no disk signature is seen in total flux. In our first polarization study of three of the rare objects known from IUE spectra to have partial Lyman edges at the systemic redshifts, we found low polarizations longward of the edges, so we did not confirm the Laor \\ea prediction. We surprisingly did find high polarization {\\it shortward} of the edge in a couple of objects, contrary to the accretion disk predictions of Laor \\etal, but qualitatively explicable by effects found in the more detailed calculations of Blaes \\& Agol (1996); see also Agol \\& Blaes (1996). The previous studies of intermediate redshift quasars (Koratkar \\ea 1995 and Impey \\ea 1995), included 4 objects, one of which shows only a marginal detection of polarization, while the remaining three objects show significant polarization ($>$ a few percent) shortward of the Lyman edge. A study of three high redshift objects observed from the ground failed to show any polarization changes at the edge position; most had tight limits on the polarization longward of the edge, but noisy data shortward of the Lyman edge (Antonucci \\ea 1996). PG 1630+377 is the only object yet observed that can be studied in any detail (Koratkar \\ea 1995; Paper I). In this object $P$ rises rapidly shortward of the edge, reaching 20\\% by 1600 {\\AA} (650 {\\AA} rest wavelength). The Ly$\\alpha$ emission line also shows a high (7.3\\%) polarization at the same position angle. Antonucci \\ea (1996) discuss polarization observations and other constraints on disk models in some detail. Based on the small number of QSOs observed in the UV in polarization, at the time of paper I, we could say little about whether high polarization shortward of the Lyman edge is common. Hence, we have significantly expanded the UV polarization database by observing six bright, $z > 1$ QSOs both below and above the Lyman edge. In this paper we discuss these new spectropolarimetric ultraviolet observations from the {\\it Hubble Space Telescope} Faint Object Spectrograph ({\\it HST/FOS}). A difference with respect to our previous study, however, is that only two of the new objects were suspected to have partial edges in total flux at the systemic redshift. The rest were simply selected because they show significant flux at short wavelengths. ", "conclusions": "Of the six QSOs identified by Koratkar, Kinney \\& Bohlin (1992) as candidate targets which have partial Lyman edges consistent with edges from simple thin accretion disks, we now have spectropolarimetric observations for five QSOs. We showed in section 3.2 that one of these candidates, 0743$-$673 no longer qualifies as a partial Lyman edge object. There are only 13 high and intermediate redshift QSOs which have spectropolarimetry observations shortward of the Lyman edge region. These objects come from this paper, Koratkar \\ea (1995), Impey \\ea (1995), and Antonucci \\ea (1996). At this point any detailed statistical tests of polarization distributions are certainly not warranted given the inhomogeneous selection criteria and data quality, and the highly model-dependent predictions. Yet, Lyman edge spectropolarimetry results can be summarized as follows: \\begin{itemize} \\item{}Of the 13 objects only three objects (PG 1630+377, PG 1338+416 and PG 1222+228 from paper I and Impey \\ea 1995) show significant polarization at wavelengths shorter than 912\\AA\\/ (Lyman edge). To these three objects we can add one more marginal detection (PKS 0405$-$123 from paper I). All 13 objects in the sample show a polarization signature which is inconsistent with any simple accretion disk model. Furthermore, $\\sim$30\\% of the sample show a rise in polarization shortward of the Lyman edge. This observed rise in polarization is qualitatively consistent with the disk models of Blaes \\& Agol (1996). A number of different interpretations of the UV signature have been given by Lee \\& Blandford (1997), Shields, Wobus \\& Husfeld (1997) and by us in paper I. We urge the interested reader to consult those papers for more details. \\item{}There are a total of five objects (PG 0117+213 from the present sample, PG1630+377, PG 1338+416, and PKS 0405$-$123 from paper I, and 0014+813 from Antonucci \\ea 1996) which show candidate partial Lyman absorption edges due to accretion disks at the systemic redshift in total flux. Of these five objects, two (PG 1630+377, PG 1338+416) have sufficient signal-to-noise at rest wavelengths of $\\leq$750\\AA\\/, and show significant polarization shortward of the Lyman edge (at least a few percent, detected at four sigma or greater significance). If the rise in UV polarization seen in PG 1630+377 is characteristic of objects with partial Lyman edges we need to observe rest wavelengths as short as $\\sim$700\\AA. The effective shortest rest wavelength observed in PG 0117+213 and PKS 0405$-$123 is $\\sim$800\\AA\\/. Thus in these objects we could have missed the rise in polarization, although we do have a marginal detection for PKS 0405$-$123. 0014+813 does not have sufficient signal-to-noise shortward of the Lyman edge. To summarize, of the objects that show candidate partial Lyman absorption edges in total flux, $\\sim$40\\% show polarization shortward of the Lyman edge. \\item{}We have eight objects in the sample of 13 that do not show a partial Lyman edge feature in total flux. Only one object out of these, PG 1222+228, from Impey \\ea (1995) shows significant polarization shortward of the Lyman edge ($P$ = 4.6$\\pm$0.9\\%). The rest show no detection of polarization in the 912\\AA\\/ spectral region. The linear polarization upper limits shortward of the Lyman edge in the remaining objects is \\ltsima 4\\%. \\item{}PG 1630+377 is the best studied object (see Paper I), and shows UV polarization reaching 20\\% at 650\\AA\\/ rest wavelength. Such high degree of polarization is rare in (non-blazar) QSOs. \\end{itemize} Perhaps the simplest result of the current data is that it strengthens the conclusion that quasars are generally not polarized significantly just longward of the Lyman edge at $\\sim$1000\\AA\\/. This paper, Koratkar \\ea (1995), Impey \\ea (1995), and Antonucci \\ea (1996) together present good observations of about 20 objects in the region just longward of the Lyman edge, all with low UV polarization (\\ltsima 1.5\\%). Further, there is no significant statistical wavelength dependence to the polarization as predicted by the models of Laor et al (1990). From these observations we conclude that simple plane-parallel atmospheres with scattering-dominated opacity are not significant sources of UV flux in quasars. Recapitulating the previous discussions here briefly, we note that models from the 1980s generally assumed that AGN accretion disks are powered by viscous dissipation below the atmospheres, and that the atmospheric opacities are dominated by electron scattering opacity in the annuli that produce the rest optical and UV. This results in 0\\% to 11.7\\% polarization (Chandrasekhar 1960), depending on inclination, and in a direction perpendicular to the symmetry axis of the disks. Pioneering optical polarimetry observations showed much smaller polarizations, which are {\\it parallel} to the axes when the latter could be inferred from a radio jet position angle (Stockman, Angel \\& Miley 1979; Antonucci 1988). To explain the low observed polarization, subsequent models by Laor et al (1990) suggested that electron scattering was only important in the $\\sim$1000-2000 \\AA\\/ range, with the Lyman continuum and free-free absorption opacity dominating shortwards and longwards of that interval respectively. Other more detailed calculations revealed that a lower fraction of absorption opacity was required to reduce the predicted polarization than was assumed by Laor et al; and that under rather special circumstances a large polarization {\\it parallel} to the disk axis could be produced shortward of the Lyman edge (Blaes and Agol 1996, and references therein). The observed rise in UV polarization shortward of the Lyman edge has been interpreted both in the context of accretion disk models and non-disk related models. Here we do not further discuss the polarization and depolarization mechanisms discussed in detail in Paper I. An additional key complication in the accretion disk models, is that AGN variability data require that the disk atmosphere is actually illuminated from above, (e.g. Antonucci 1988, Sincell and Krolik 1996), perhaps producing a strong polarization which cannot be calculated rigorously without specification of the illumation geometry." }, "9804/gr-qc9804041_arXiv.txt": { "abstract": "{The mass function of primordial black holes created through the near-critical gravitational collapse is calculated in a manner fairly independent of the statistical distribution of underlying density fluctuation, assuming that it has a sharp peak on a specific scale. Comparing it with various cosmological constraints on their mass spectrum, some newly excluded range is found in the volume fraction of the region collapsing into black holes as a function of the horizon mass. } \\pacs{PACS Numbers: 04.70.Bw, 04.70Dy, 98.80.Cq } ", "introduction": " ", "conclusions": "" }, "9804/astro-ph9804140_arXiv.txt": { "abstract": "We investigate whether models based on the assumption that jets in quasars are powered by rotating black holes can explain the observed radio dichotomy of quasars. We show that in terms of the ``spin paradigm'' models, radio-loud quasars could be objects in which the black hole's rotation rate corresponds to an equilibrium between spin-up by accretion and spin-down by the Blandford-Znajek mechanism. Radio-quiet quasars could be hosting black holes with an average spin much smaller than the equilibrium one. We discuss possible accretion scenarios which can lead to such a bimodal distribution of black hole spins. ", "introduction": "Quasars are characterized not only by an intense radiation from the central engine, but also by jets which power large scale radio structures. The ratio of radio luminosity of these structures to the optical luminosity of the central sources shows a bimodal distribution, with only $\\sim 10$ \\% of quasars belonging to the radio-loud category (see, e.g., Kellerman et al. 1989; Hooper et al. 1995; Falcke, Sherwood, \\& Patnaik 1996; Bischof \\& Becker 1997). Radio-loud quasars have never been found in spiral galaxies, while the hosts of radio-quiet quasars are either spiral or elliptical galaxies (Taylor et al. 1996; Kukula et al. 1998). On the other hand, both radio-quiet and radio-loud quasars have almost identical average IR-optical-UV spectra (Francis et al. 1993; Zheng et al. 1997), which suggests similar accretion conditions in these two samples of objects. These properties seem to support models which are based on the idea that jets are powered by rotating black holes. However, in order to explain the observed radio loudness bimodality in terms of a bimodal distribution of black hole spins, one has to assume that the population of supermassive black holes is dominated by very low-spin black holes, and one should understand why radio-loud quasars are always hosted by elliptical galaxies. This issue has been addressed by Wilson \\& Colbert \\shortcite{wc95}, who proposed that high-spin black holes exist only if formed by coalescence of two black holes. Such coalescences would take place mostly in a denser galactic environment (groups, clusters), which are much more populated by elliptical galaxies than the field regions. However, the basic assumption of this scenario, that black holes which do not undergo coalescence rotate very slowly -- despite the angular momentum gained from accretion discs -- must be verified. In other words, one should verify that there exists a mechanism which could keep black holes at low rotation rates despite of the accreted angular momentum. This problem was investigated by Moderski \\& Sikora \\shortcite{ms96a}. They showed that the Blandford-Znajek (B-Z) mechanism, which extracts rotational energy from a black hole \\cite{bz77}, is not efficient enough to counteract the spinning up of a black hole by the gas accreted from standard $\\alpha$-discs \\cite{ss73}. Low-spin equilibrium states are possible only for very low accretion rates but the time required to established such an equilibrium is longer than the age of the Universe. This is because the power extracted from the black hole is proportional to the square of intensity of the magnetic field which threads the black hole. The field intensity is limited by the pressure in the accretion flow and, therefore, is very low for very low accretion rates. Hence, once a black hole gets a high spin, it will be rotating fast forever. Low-spin equilibrium solutions for high accretion rates are possible only for the so called $\\beta$-discs \\cite{ms97}. In such discs the viscous stress driving accretion is proportional to the gas pressure only (and not to the total pressure as in the case of $\\alpha$ discs). Since the relation between the viscous stress and the pressure in radiation pressure dominated accretion discs is unknown, such discs are a reasonable alternative to $\\alpha$-discs. For high accretion rates the gas pressure is by several orders of magnitude smaller than the radiation pressure \\cite{ss73}. This implies that, for a given accretion rate, $\\beta$-discs are much denser than $\\alpha$-discs \\cite{sc81} and, therefore, can confine much stronger black hole magnetic fields. Since the power extracted from a black hole via the B-Z mechanism scales with the square of the magnetic field intensity, equilibrium spins for $\\beta$-discs are smaller than for $\\alpha$-discs. This, of course, does not mean that the power extracted from a black hole with a $\\beta$-disc must be smaller than the power extracted from a black hole with an $\\alpha$-disc, because in the expression for the B-Z power a lower equilibrium spin is compensated by a higher intensity of the magnetic field \\cite{ms97}. Thus, low equilibrium spin black holes can represent radio-quiet quasars only, if the conversion of the extracted black hole energy into jet energy is very inefficient. According to the Wilson-Colbert scenario powerful jets would then exist only in objects where a coalescence of two black holes leads to spins much higher than the equilibrium one. However, powers extracted from black holes with spins much larger than the equilibrium one are so large, that such black holes would be spun-down on time scales two orders of magnitude shorter than the typical lifetime, $\\sim 10^8$ years, of radio quasars (see, e.g., Leahy, Muxlow \\& Stephens 1989). Powering jets in radio-loud quasars by a black hole can last as long as $10^8$ years, only if losses of angular momentum due to the B-Z mechanism are compensated by gains of angular momentum from an accretion disc. This would give a correlation between radio and optical luminosities, in accordance with observations \\cite{ser97}. Then, however, one would have to explain why the majority of super-massive black holes have spin values much lower than the equilibrium one -- the condition for the existence of radio-quiet quasars. As suggested by Moderski, Sikora \\& Lasota \\shortcite{msl97}, black holes in most objects could be forced to rotate slowly by multi-accretion events with random orientations of the angular momentum vector. In such a scenario, quasars which become radio-loud are only those which undergo major accretion events, induced, i.e., by a merger of two big galaxies. Following this, a black hole can easily double its mass and reach an equilibrium state. In this paper we explore this possibility and derive conditions the model must satisfy in order to explain the observed radio properties of quasars. This paper is organized as follows. In Section 2 we present equilibrium spin solutions, obtained for a variety of accretion disc models. In Section 3, we discuss multi-event accretion scenarios. In Section 4 we use our results to derive conditions which must be satisfied by quasar evolution models in order to obtain bimodal distribution of black hole spins. ", "conclusions": "Radio dichotomy of quasars was discovered many years ago (Strittmatter et al. 1980; Kellerman et al. 1989), but is still waiting for a theoretical explanation. As for now, the consensus concerns only one aspect of the problem: it is clear that jets in quasars must be formed near a supermassive black hole. This follows from the energetics of quasar jets, since no other known sources could power jets at a rate reaching $10^{46}$ ergs s$^{-1}$ for millions of years (Rawlings \\& Saunders 1991; Leahy et al. 1989). Independent argument for the formation of extra galactic jets in the vicinity of supermassive black holes is provided by direct VLBI observations of nearby radio galaxies. In particular, in 3C 274 (M87) the jet is seen down to $10^{16}$ cm from the center \\cite{jb95}, which corresponds to $100$ gravitational radii for the $3 \\times 10^9 M_{\\odot}$ \\cite{har94} central black hole. There, very deep in the gravitational potential well, jets could be powered either by the innermost parts of an accretion disc (Blandford \\& Payne 1982; Park \\& Vishniac 1994; Contopoulos 1995; Begelman 1995) or by a rotating black hole (Blandford \\& Znajek 1977; Rees et al. 1982). However, no jet production model can be successful, if it fails to explain why only a small fraction of quasars is radio loud and why radio-loudness has a bimodal distribution. In terms of the $R=F_r/F_o$ ratio, where $F_r$ and $F_o$ are the monochromatic fluxes measured at frequencies $\\sim 10^{10}$ Hz and $\\sim 10^{15}$ Hz, respectively, radio-quiet quasars cluster around $R \\sim 0.3$ and radio-loud quasars cluster around $R \\sim 300$ (Kellermann et al. 1989; Falcke, Sherwood \\& Patnaik 1996). Thus, the average radio-loudness of the two quasar populations differs by a factor $10^3$ and this number, together with typical radio luminosities of radio loud quasars, $L_r \\sim 10^{45}$ ergs s$^{-1}$, provides the basic quantitative conditions which should be satisfied by any unified model of quasars. These conditions, together with our results discussed in the two previous sections are used below to test spin based models of a jet activity in quasars. The predictions of such models should also satisfy such observationally established trends, as \\noindent - radio-loud quasars avoid disc-galaxies and have UV-luminosities $ \\ge 10^{46}$ ergs s$^{-1}$; \\noindent - radio-quiet quasars are present both in spiral and elliptical galaxies \\cite{tay96} and their radio properties do not depend on the galaxy morphology \\cite{kuk98}; \\noindent - radio properties of radio-quiet quasars suggest that they are, like in radio loud quasars, related to the jet production by a central engine. Assuming that the efficiency of conversion of jet energy into radio emission is $\\sim 10$\\%, the typical jet in radio-loud quasars should have $P~\\sim~10^{46}$ ergs s$^{-1}$. Similar jet powers are deduced by calculating the total energy content of extended radio sources and dividing it by the age of the source \\cite{lea89}, or from energetics of $\\gamma$-ray production in sub-parsec jets (see, e.g., Sikora 1997). Largest powers which can be extracted from rotating black holes are given by equation (\\ref{power}). For $A=1$ and $B_{\\perp} = 8\\pi p_{tot}$ we obtain $P_{max} \\sim 3 \\times 10^{44} M_8^2 p_{tot,8}$ ergs s$^{-1}$, where $p_{tot,8} = p_{tot}/10^8$dyne cm$^{-2}$ and $M_8 = M/10^8 M_{\\odot}$. One can see from Figure~\\ref{pmax} that for high accretion rates ($\\dot m > 0.1$, say) and black hole masses $\\sim M=10^9 M_{\\odot}$, a pressure $\\ge 10^8$ dyne cm$^{-2}$ is provided by $\\alpha$-discs with $\\alpha \\le 0.1$, and by all $\\beta$-discs. Thus, the B-Z mechanism is efficient enough to power jets in radio-loud quasars, provided the black hole magnetic field is supported by the total disc pressure. If the latter is not true and, as Ghosh and Abramowicz \\shortcite{ga97} argued, the energy density of the black hole magnetic field cannot exceed the energy density of the maximum magnetic field in a disc, then the maximum pressure of the black hole's magnetic field is numerically equal to the total pressure in an $\\alpha=1$ disc. In this case black hole masses $\\sim 3 \\times 10^9 M_{\\odot}$ are required in order to get $P \\sim 10^{46}$ ergs s$^{-1}$. The case of M87 seems to prove that such black holes are not necessarily exceptional \\cite{har94}. However, one should note here, that the question of the diffusion of an external magnetic field into an accretion disc is still open (see, e.g., Wang 1995; Bardou \\& Heyvaerts 1996). Assuming, as before, that the fraction of the jet energy converted into radiation is 10\\%, and that the bolometric corrections for jet radiation at $\\sim 10^{10}$Hz and for accretion disc radiation at $\\sim 10^{15}$Hz are of the same order, we obtain that $P/L_d \\sim 10 L_r/L_o \\sim 10 F_r{\\nu}_r/F_o{\\nu}_o \\sim 10^{-4} R$. Thus, for radio-loud quasars models should predict $P/L_d \\sim 0.1$, while radio-quiet quasars should cluster around $P/L_d \\sim 10^{-4}$. As is seen from Figures~2 and 3, $P/L_d \\sim 0.1$ nicely corresponds to black hole equilibrium spin solutions for all but $\\alpha > 0.1$ disc models. One can also check, that there are no equilibrium spin solutions which would correspond to radio-loudness of radio-quiet objects. For them $A < 0.03$ is required, provided that radio luminosity scales linearly with $P$. A population of such low spin black holes can exist only if black holes are born with very low spin and then accrete very little (Moderski, Sikora \\& Lasota 1997), or if black hole evolution is determined by multi-accretion events with random angular momenta. As one can deduce from results presented in Figure~7, hundreds of accretion events per object are required in order to have more than $90$ \\% of black holes with $A < 0.1$ at any given moment. This is too much to be obtained by accretion events induced by capture of dwarf galaxies, but can be achieved by accretion of molecular clouds. Molecular cloud accretion events were recently proposed by Sanders \\shortcite{san98} to explain some properties of Sgr A$^*$ and other AGNs. This scenario is supported by the random orientation of central engines vs. the orientation of galactic discs, as deduced from observations of ``UV'' cones (Wilson \\& Tsvetanov 1994; McLeod \\& Rieke 1995) and radio axis \\cite{ckp98} in Seyfert galaxies. Here we should note, that in our simplified treatment of the multi-accretion scenario (Section~3), we didn't take into account the coupling between the spin of the black hole and the orbital angular momentum of the approaching molecular clouds. Such coupling supposedly leads to random wondering of the black hole spin vector. One can now speculate that changes of orientation of the black hole spin could be interrupted and the black hole could be spun-up to very high spins following a merger process. This process could induce a massive and long lasting accretion event. If during such an event the accretion proceeds from a fixed plane and at least doubles the black hole mass, the black hole spin reaches the equilibrium spin and the object becomes a typical radio-loud quasar \\cite{msl97}. Since mergers happen mostly in groups and clusters of galaxies, where the population of galaxies is dominated by ellipticals, this could explain why radio loud quasars avoid spiral galaxies. Observational arguments for such a scenario are exactly the same as those used by Wilson and Colbert \\shortcite{wc95}. The only difference is, that they postulated formation of high spin black holes via coalescence of two supermassive black holes, while in our scenario high spins result from an accretion process. Note, however, that a coalescence of two black holes, if it happens, does not have to affect much our scenario. If the coalescence involves two black with very different masses, the final spin will be determined by the accretion process, otherwise both processes lead to similar spins. What are the perspectives for an observational test of the assumption that radio-quiet objects have low spins? A possibility to measure the spin of supermassive black holes is provided by the detailed studies of profiles of the X-ray fluorescent iron line produced in the surface layer of the innermost parts of accretion discs. Such lines are detected in many Seyfert galaxies, which represent the low luminosity branch of radio-quiet quasars. For at least one of such objects the line profile was claimed to be consistent with the kinematics given by the rotation of a disc around a black hole in fast rotation \\cite{iwa96}. However, as demonstrated by Reynolds and Begelman \\shortcite{rb97}, similar line profiles can be produced around non-rotating black holes, provided that a large part of the line emission comes from below the marginally stable orbit. Therefore, much more detailed theoretical models and sensitive observations are required to get conclusive diagnostics from this type of investigations. The remarkable discovery of relativistic jets in several Galactic X-ray sources (cf. Mirabel \\& Rodriguez 1994; Hjellming \\& Rupen 1995; Newell, Spencer \\& Garrett 1997) suggests that the radio-dichotomy exists for Galactic compact objects as well. As was argued recently by Zhang, Cui \\& Chen \\shortcite{zcc97}, jet activity in these sources can also be conditioned by the value of the black hole spin." }, "9804/astro-ph9804154_arXiv.txt": { "abstract": "We study the interpretation of the mean surface density of stellar companions as a function of separation (or, equivalently, the two point correlation function of stars) in star-forming regions. First, we consider the form of the functions for various simple stellar distributions (binaries, global density profiles, clusters, and fractals) and the effects of survey boundaries. Following this, we study the dependencies of the separation at which a transition from the binary to the large-scale clustering regime occurs. Larson \\shortcite{Larson95} found that the mean surface density of companions follows different power-law functions of separation in the two regimes. He identified the transition separation with the typical Jeans length in the molecular cloud. However, we show that this is valid only for special cases. In general, the transition separation depends on the volume density of stars, the depth of the star-forming region, the volume-filling nature of the stellar distribution, and on the parameters of the binaries. Furthermore, the transition separation evolves with time. We also note that in young star-forming regions, binaries with separations greater than the transition separation may exist, while in older unbound clusters which have expanded significantly, the transition contains a record of the stellar density when the stars formed. We then apply these results to the Taurus-Auriga, Ophiuchus, and Orion Trapezium star-forming regions. We find that while the transition separation in the Taurus-Auriga star-forming region may indicate a typical Jeans length, this is not true of the Orion Trapezium Cluster. We caution against over-interpreting the mean surface density of stellar companions; while Larson showed that Taurus-Auriga is consistent with the stars having a fractal large-scale distribution we show that Taurus-Auriga is also consistent with stars being grouped in non-hierarchical clusters. We also argue that to make a meaningful study of the stellar distribution in a star-forming region requires a relatively complete stellar survey over a large area. Such a survey does not currently exist for Ophiuchus. Finally, we show that there is no evidence for sub-clustering or fractal structure in the stars of the Orion Trapezium Cluster. This is consistent with the fact that, if such structure were present when the stars formed, it would have been erased by the current age of the cluster due to the stellar velocity dispersion. ", "introduction": "\\label{introduction} Stars generally do not form in isolation. Instead, on small scales, they frequently form as members of bound binary or higher-order multiple systems (e.g. Duquennoy \\& Mayor 1991; Mayor et al. 1992; Fischer \\& Marcy 1992; Ghez, Neugebauer, \\& Matthews 1993; Leinert et al. 1993, Simon et al. 1995), while on larger scales they are often members of associations or clusters of stars (e.g. Gomez et al. 1993; Lada, Strom, \\& Myers 1993; Zinnecker, McCaughrean, \\& Wilking 1993). Studying the clustering properties of stars on different length scales may help to determine what processes are involved in their formation. Gomez et al. \\shortcite{GHKH93} found that the pre-main-sequence stars in the Taurus-Auriga molecular cloud are not randomly distributed, but instead are in small associations of $\\sim 15$ stellar systems within radii of $\\sim 0.5-1.1$ pc. As one method of analysing the spatial distribution of stars, Gomez et al. determined the two-point angular correlation function and found that it could be represented by a single power-law over separations from $0.005$ to $5$ pc, implying that stars are clustered self-similarly. However, they also found weak evidence that two-point angular correlation function may be better represented by two different power laws with a break at $\\approx 0.05$ pc. Using data from searches for binary companions to pre-main-sequence stars in the Taurus-Auriga molecular cloud, Larson \\shortcite{Larson95} extended the two-point angular correlation function to smaller separations than Gomez et al. \\shortcite{GHKH93} and demonstrated that, indeed, there is a break at $\\approx 0.04$ pc. Rather than using the standard two-point angular correlation function, Larson used the closely-related mean surface density of companions (MSDC) (see Section \\ref{MCSD}). The MSDC has the advantage that no normalisation is required, whereas the two point correlation function must be normalised by the average density in the survey area which can be difficult to determine if the stars are clustered. Larson \\shortcite{Larson95} found that, for stars in the Taurus-Auriga molecular cloud, the MSDC has a power-law slope of $\\approx -0.6$ on large scales, but steepens below $\\approx 0.04$ pc with a slope of $\\approx -2$ on small scales. The fact that a break occurs indicates that a single scale-free process is not responsible for the formation of stars on both scales. The power-law slope of $\\approx -0.6$ on large scales is due to the clustering of stellar systems that Gomez et al. \\shortcite{GHKH93} studied. Furthermore, Larson pointed out that a power-law slope of $-0.6$ means that the number of stars within an angular distance $\\theta$ of an average star increases as $\\theta^{1.4}$ and, thus, the distribution of stars on this scale can be described as a fractal point distribution with dimension 1.4. Larson identified the power-law slope of $-2$ for small angular separations with the distribution of binary separations, since stellar pairs closer than $0.04$ pc in Taurus-Auriga are typically mutually bound. However, the power-law slope of $\\approx -2$ is not due to a fractal distribution. Rather, it results from the fact that the frequency distribution of binary separations is roughly uniform in log-separation \\cite{DuqMay91}. Finally, Larson noted that the length scale of $\\approx 0.04$ pc is essentially equal to the typical Jeans length in the Taurus-Auriga molecular cloud. Thus, Larson associated the location of the break in the MSDC with the Jeans length, speculating that companions with separations smaller than this formed due to the fragmentation of a single collapsing molecular cloud core, while on larger scales stars are grouped self-similarly due to hierarchical structure in the progenitor molecular clouds. Following Larson's analysis of the Taurus-Auriga star-forming region (SFR), Simon \\shortcite{Simon97} considered the spatial distribution of stars in the Ophiuchus and Orion Trapezium regions. As with Taurus-Auriga, a break was found in the MSDC for each region. On small scales, both Ophiuchus and the Orion Trapezium could be fit by power laws with slopes of $\\approx -2$. On large scales, flatter power laws were required of $-0.5\\pm0.2$ for Ophiuchus and $-0.2\\pm0.2$ for the Orion Trapezium. However, the break between the two regimes was found to occur at $\\approx 400$ AU for the Orion Trapezium and $\\approx 5000$ AU for Ophiuchus, compared to $\\approx 10000$ AU (taking the mean of Simon's and Larson's results) for Taurus-Auriga. Simon concluded that all three SFRs had similar distributions of binary separations and similar fractal structure on large scales, but that the location of the break seemed to depend not only on the Jeans length, but also on the stellar density of the SFR. Finally, Nakajima et al. \\shortcite{NTHN98} considered the MSDC of stars in the Orion, Ophiuchus, Chamaeleon, Vela, and Lupus star-forming regions. Again, for those regions where the survey data extends to small enough separations, they find a break in the MSDC with a power-law slope of $\\approx -2$ on small scales and power-law slopes ranging from $-0.15$ to $-0.82$ on large scales. The location of the break was also found to vary from a minimum of $\\approx 1000$ AU to a maximum of $\\approx 30000$ AU. Nakajima et al. also considered the nearest-neighbour distributions for each of the regions and found that when the nearest-neighbour distribution could be fit well by a Poisson distribution, the MSDC had a power-law index close to zero on large-scales, while when the nearest-neighbour distribution was broader than the Poisson distribution, the MSDC had a large, negative power-law index. They interpreted this as evidence that the MSDC may indicate a star formation history in the region rather than the presence of self-similar spatial structure; if the stars have a range of ages, the older stars typically will be more dispersed than the younger stars resulting a spread in the distribution of separations of nearest neighbours and a range of stellar surface density which provides the slope of the large-scale MSDC. Motivated by these papers, we make a careful study of the interpretation of the mean surface density of companions (MSDC) of star-forming regions. Amongst other goals, we wish to determine the relationship of the break between the binary and large-scale regimes to the Jeans length and the stellar density in star-forming regions. We also want to determine how sensitive the MSDC is to detecting sub-structure in a stellar distribution and, when detected, what can be said about the form of the sub-structure (e.g. whether the sub-structure is self-similar or not) and how robust the result is. In Section \\ref{MCSD} we consider the calculation of the MSDC function, handling of survey boundaries, and the results for simple stellar distributions (binaries, global density profiles, clusters, and fractals). In Section \\ref{posbreak} we derive the dependencies of the break between the binary and large-scale regimes, and show that the separation at which the break occurs can only be identified with the Jeans length in special cases. We also indicate how the MSDC of SFRs is expected to evolve with time. Based on these results, we reconsider the Taurus-Auriga, Ophiuchus, and Orion Trapezium star-forming regions in Section \\ref{application}. Finally, we present our conclusions in Section \\ref{conclusions}. ", "conclusions": "\\label{conclusions} We have studied the interpretation of the mean surface density of companions (MSDC) as a function of separation $\\Sigma_{\\rm com}(\\theta)$ in star-forming regions. We have shown how the power-law slope of $\\approx -2$ for binaries is due their flat distribution of periods in the logarithm of separation, and have considered the MSDC of various global density profiles, sub-clusters and self-similar distributions. We emphasise that simply because a power-law slope can be fit to a particular MSDC, it does not mean that the stellar distribution is self-similar or fractal. We have also demonstrated the effects of survey boundaries on the calculation of $\\Sigma_{\\rm com}(\\theta)$. Several methods of attempting to avoid boundary effects were considered, all of which provide a full correction in the case that there is no large-scale stellar density gradient across boundaries, but none of which give a perfect correction when there are such large-scale gradients. Of these, we recommend Method 5, since it allows the maximum range of separations to be studied, does not discard any information, and is simple to use for surveys with irregular boundaries. Even in the case of a uniform stellar distribution, the improper handing of boundaries results in the $\\Sigma_{\\rm com}(\\theta)$ having a significant slope for separations greater than $\\approx 1/50$ of the survey area's dimensions (i.e.~using Method 1). Larson \\shortcite{Larson95} associated the separation at which a break in $\\Sigma_{\\rm com}(\\theta)$ between the binary regime and the large-scale regime occurs with the Jeans length in the Taurus-Auriga star-forming region (SFR). However, we show this transition separation may only be associated with the Jeans length in special cases, and that the transition separation does not necessarily give the maximum binary separation. In general, the break occurs at the separation where the mean surface density of {\\em binary} companions is equal to the mean surface density of {\\em non-binary} companions (the latter of which may be physically close, or simply chance projections). Thus, typically, the break occurs at smaller separations for SFRs with higher stellar surface densities (as observed by Simon \\shortcite{Simon97} and Nakajima et al. \\shortcite{NTHN98}). In turn, the surface density of non-binary companions depends on the parameters of the binaries, the volume density of stars, the volume-filling factor of the stellar distribution and, in general, the depth of the star-forming region. The transition separation between the binary and the large-scale regimes also evolves with time. Due to a stellar velocity dispersion, initial structure is erased and the surface density of stars in an unbound region generally decreases. This effect begins at the smallest scales, extends to larger scales with time, and results in the transition separation increasing with time. Finally, the transition between the binary and the large-scale regimes may allow a truncation of binaries at large separations to be detected, especially in old clusters that were much denser when the stars were formed and have since expanded. In such cases, this provides a record of the stellar density when the stars first formed. In summary, the transition separation may be associated with the Jeans length only if the star-forming region is young enough that initial structure has not been erased, and if the SFR is `optically thin' in the sense that projection effects due to the depth of the SFR do not affect the transition separation. The latter is true if the volume-filling factor of the SFR is low (e.g. the SFR is composed of widely separated clusters consisting of only a few stars ($\\sim 10$), or if the stars have a fractal distribution with dimension $\\simless 1.5$). This is the case for the Taurus-Auriga SFR, which explains the good agreement between between the transition separation and the Jeans length found by Larson \\shortcite{Larson95}, but it is not the case for the Orion Trapezium Cluster. It is important when studying the large-scale spatial distributions of star-forming regions to obtain the most complete sample of stars over the largest area possible. The lack of such data for the Ophiuchus SFR makes an attempt to study its large-scale spatial distribution of little use at this time. For the Taurus-Auriga and Orion Trapezium SFRs, the current data makes a meaningful study of their large-scale stellar distribution possible. For the Taurus-Auriga SFR, Larson \\shortcite{Larson95} fit the large-scale MSDC with a power-law slope that implied a fractal stellar distribution. However, this is not the only possible interpretation; the data can be equally well fit by assuming the stars are formed primarily in randomly-distributed clusters of stars. For the Orion Trapezium SFR, we find that the MSDC is consistent with the stars simply being distributed according to a surface density that decreases with radius; there is no evidence for sub-structure (either fractal or sub-clusters) in the stellar distribution. We also demonstrate how upper limits can be placed on how much sub-clustering is present, and note the the sensitivity of the MSDC to detecting sub-structure appears to be slightly less than that of the human eye. The results for the Orion Trapezium SFR are consistent with the fact that if structure were present when the stars formed, it would have been erased by the current time due to the stellar velocity dispersion. Binaries in the Taurus-Auriga and Orion Trapezium SFR are roughly consistent with an MSDC with a power-law slope of $\\approx -2$. However, we point out that comparing power-law indices derived from the slope of the MSDC in the binary regime is not the best way to compare the distribution of binary separations between stellar populations since any structure or deviation from a true power-law may easily be missed. In the centre of the Orion Trapezium SFR, we find very weak evidence that there may be a deficit of binaries with separations $\\simgreat 500$\\,AU\\@. Such a deficit may be caused by the disruption of wide binaries by single-binary star encounters. Finally, in view of our studies of the Taurus-Auriga and Orion Trapezium SFRs, we emphasise caution when interpreting the MSDC. Rather than attempting to characterise star-forming regions simply by fitting power-laws to $\\Sigma_{\\rm com}(\\theta)$, it is more instructive to also consider the global stellar distribution (e.g. a radial surface density profile) and to compare the MSDC to those of model stellar distributions to determine the robustness of any conclusions. Alternatively, rather than just considering the MSDC (or, equivalently, the two-point correlation function), correlation functions of higher order (three and four-point correlation functions) and/or the nearest-neighbour distribution can be used to differentiate between non-hierarchical and hierarchical structure. The use of higher-order correlation functions is common in studying the large-scale structure of the universe \\cite{Peebles80}. The nearest-neighbour distribution has been used by Nakajima et al. \\shortcite{NTHN98} to argue that the power-law slope of an MSDC on large scales may indicate a stellar age spread rather than the presence of hierarchical structure. However, while an age spread does help explain their results, we argue that their results do not exclude the possibility that stars form in hierarchical structures. More work is required on this topic." }, "9804/astro-ph9804012_arXiv.txt": { "abstract": "The diffuse X-ray emission from the thin disk surrounding the Galactic mid-plane (the so-called Galactic ridge) was measured with {\\it RXTE} PCA in order to determine the spatial extent, spectral nature, and origin of the emission. Spatial examination of the diffuse emission in the central $30^\\circ$ of the plane in Galactic longitude reveals the presence of two components: a thin disk of full width $ \\lesssim 0^\\circ \\!.5$ centered roughly on the Galactic mid-plane, and a broad component which can be approximated as a Gaussian distribution with FWHM of about $ 4^\\circ$. Assuming an average distance of 16~kpc to the edge of the galaxy, a scale height of about $ 70$~pc and 500~pc is derived for the thin and broad disk components, respectively. Spectral examination of the emission clearly reveals the presence of a hard power law tail above 10~keV and an emission line from He-like iron, indicating both thermal and possibly non-thermal origins for the diffuse emission. The averaged spectrum from the ridge in the $3-35$~keV band can be modelled with a Raymond-Smith plasma component of temperature $\\sim 2-3$~keV and a power law component of photon index $\\sim 1.8$. Based on this finding, we argue that the temperature of the hot phase of the Interstellar Medium (ISM) is less than the previously reported values of $5-15$~keV. Motivated by the similarities between the characteristics of the thermal component of the Galactic ridge emission in our model and the thermal emission from supernova remnants (SNRs), we discuss the origin of the thermal emission in terms of a population of SNRs residing in the Galactic disk. We find that a SN explosion rate of less than 5 per century is adequate to power the thermal emission from the ridge. The origin of the emission in the hard X-ray band modelled by a power law remains uncertain. Possible contributions from non-thermal bremsstrahlung of cosmic ray electrons and protons, inverse Compton scattering of energetic electrons from ambient microwave, infrared, and optical photons, non-thermal emission from SNRs, and emission from discrete X-ray sources are discussed. We speculate that bremsstrahlung of accelerated electrons and protons in SNR sites can play a significant role in producing the hard tail of the spectrum. Moreover, their collisional losses can play a major role in the ionization of the ISM. ", "introduction": "A galaxy can be well described by an ecosystem. There is an intimate relationship, much like a symbiosis, between the discrete components of the galaxy such as stars, and its interstellar medium (ISM). The ISM provides the foundation for the birth of a new generation of stars, while at the same time it is enriched by the remains of the older generations and their byproducts during their life cycle. Hence, it is natural to expect that the understanding of formation and evolution of galaxies is closely related to the understanding of their ISM. The ISM of the Milky Way has been found to have many components. Magnetic fields and cosmic ray gas compose the relativistic fluid, while the gaseous phase consists of both ionized and neutral components. The hot ionized medium observable in UV and X-ray has a temperature above $\\sim 10^5$~K and is composed of hot coronal gas heated by supernova shocks. A good portion of the energy of the ISM resides in this component. The warm ionized medium (e.g. HII, planetary nebulae) is visible in ${\\rm H\\alpha}$, UV, and optical, and has a temperature as high as $10^4$~K. The neutral atomic gas appears to have both cold ($<100$~K; e.g. HI~clouds) and warm (100~K~$